\caption{\textcolor{blue}{\label{fig:ST SSPM and DSPM}}ST scaling laws. Set $N_{\mathrm{U}}=1$ when $N_{\mathrm{a}}=1$, while set $N_{\mathrm{U}}=2$ when $N_{\mathrm{a}}>1$. For system settings, set $P=23$dBm and $\tau=10$dB. Set $\alpha_{0}=4$ for SSPM, and $\alpha_{0}=2.5$, $\alpha_{1}=4$ and $R_{1}=10$m for DSPM. In (a), set $\Delta h=2$m. In (b), set $N_{\mathrm{a}}$=16. Lines and markers denote numerical and simulation results, respectively.}
\caption{{\color{blue}(Color online)} Illustration of a magnetic impurity setup for $^{173}$Yb atoms (schematic). Atoms in the ground-state \1S0 form a Fermi sea (Fermi energy$\epsilon_F$), while atoms in the excited-state \3P0 are trapped in an optical potential and form a dilute concentration of localized magnetic impurities.$\epsilon_{\e,\g}$ is the excitation energy (\ref{energy-eg}) of the \3P0 state.}
\caption{{\color{blue}(Color online)} Scattering lengths $a_S$ [blue curve] and $a_T$ [red curve] as functions of $V_0$.}
\caption{{\color{blue}(Color online)} Kondo temperature (blue curves) (\ref{TK-vs-scatt-lengths}) as a function of $V_0$ for $\epsilon_F/k_B = 100$~nK. The red segment indicates the interval $V_0<V_{\mathrm{c}}$ for which the Kondo effect is absent. Thus, the red dot denotes the critical potential $V_{{\mathrm{c}}}$ which separates the interval where Kondo effect is present from that where it is absent.}
\caption{\red{We plot the} Left-Hand-Side (LHS) (solid black curve with the characteristic branches of the $tan$ function) and the Right-Hand-Side (RHS) of Eq.~(\ref{even_eig}) (i.e. the even case) vs. $z \equiv ka/(2\pi)$ \red{to show the solutions graphically}. The RHS depends on the strength of the $\delta$-function potential, $\alpha$. In all cases the RHS is linear in $z$ with a zero intercept, and has a negative (positive) slope for the repulsive, $\alpha > 0$ (attractive, $\alpha < 0$) case. Two example cases are shown, the red dashed curve with $z_0 \equiv m\alpha a/(2\hbar^2) = 50$ (repulsive, $\alpha > 0$), and the blue dotted curve with $z_0 =-50$ (attractive, $\alpha < 0$). The solutions are shown by circles where these respective lines intercept the $tan$ function. The solutions for the odd case are trivial, and are simply given by $ka = 2\pi n$, $n=1,2,3,...$, and are indicated by the squares along the $z$-axis. }
\caption{The two lowest energy scattering states: their wave functions (top panels) and their probabilities (bottom panels) for a large value of $|\alpha| = 50 a\ \pi^2 \hbar^2/(2ma^2)$ for $\alpha > 0$ (left side) and $\alpha < 0$ (right side) as a function of position. Even wave functions are shown with a solid (red) curve, while odd wave functions are shown with a dashed (black) curve. Their probability amplitudes are barely distinguishable from one another. The left and right sides also look nearly the same, indicating that when the strength is high enough, a $\delta$-function moat is as effective \red{a barrier} as a $\delta$-function wall. Close inspection of the bottom two panels indicates that the moat actually keeps more wave function amplitude away from the central ``barrier'' region than the wall does (see Fig.~\ref{Fig4} below). The green vertical arrow is a schematic indication of the potential involved (repulsive on the left, and attractive on the right).}
\caption{\small \textbf{LHPs generalized improper flexoelastic model.} \textbf{a}, Schematics of hybrid layered compounds regarded as heterostructures L1/L2 with L1 the 3D (n=$\infty$) bulk materials, \textit{e.g.} MAPbI$_{\text{3}}$, and L2, a n=1 compound, \textit{e.g.} (BA)$_{\text{2}}$PbI$_{\text{4}}$. \textbf{b}, In-plane expansion and out-of-plane contractions of experimental lattice constants for (BA)$_{\text{2}}$(MA)$_{\text{n-1}}$Pb$_n$I$_{\text{3n+1}}$ and the L1 and L2 layers. The room-temperature structures of MAPbI$_{\text{3}}$ and (BA)$_{\text{2}}$PbI$_{\text{4}}$ serve as references for L1 and L2 structures, respectively. \textbf{c}, Same from the improper flexoelastic model (see \textcolor{orange}{Method} for details). \textbf{d}, Computed elastic energy density for the (BA)$_{\text{2}}$(MA)$_{\text{n-1}}$Pb$_{\text{n}}$I$_{\text{3n+1}}$ heterostructure. }
\caption{\small \textbf{Design of LHPs for photovoltaics and optoelectronics.} \textbf{a}, Lattice mismatch between various monolayered A'$_{\text{2}}$PbI$_{\text{4}}$ perovskites (n=1) and MAPbI$_{\text{3}}$ (\textit{I4cm}; n=$\infty$). All data are taken from X-ray structures resolved at room-temperature. Names for organic compounds and corresponding references are given \textcolor{orange}{Table \small\bf S4}. \textbf{b}, Computed elastic energy density for heterostructures built with MAPbI$_{\text{3}}$ and (BA)$_{\text{2}}$PbI$_{\text{4}}$ (grey line), (C$_{\text{9}}$H$_{\text{19}}$NH$_{\text{3}}$)$_{\text{2}}$PbI$_{\text{4}}$ (NoA, blue line), and (4Cl-C$_{\text{6}}$H$_{\text{4}}$NH$_{\text{3}}$)$_{\text{2}}$PbI$_{\text{4}}$ (4Cl-PhA, red line). }
\caption{Nonlinear frequency response of a clamped-clamped cross beam structure for four different levels of excitation (\textcolor{gray}{$\boldsymbol{-}$}). The response of the structure is shown at the excited DOF (co-located). The green curves (\textcolor{green}{$\boldsymbol{--}$}) trace out the loci of periodic responses that are in phase quadrature with the excitation, and the thick black curves (\textcolor{black}{$\boldsymbol{-}$}) show the NNMs of the underlying conservative system.}
\caption{NNMs of the beam structure. NNM 1: amplitude of modes 1 (\textcolor{black}{$\boldsymbol{-}$}) and 2 (\textcolor{black}{$\boldsymbol{--}$}). NNM 2: amplitude of modes 1 (\textcolor{blue}{$\boldsymbol{-}$}) and 2 (\textcolor{blue}{$\boldsymbol{--}$}). (b-g) Deformation of the structure at the locations reported on the frequency-amplitude curves.}
\caption{(a) Comparison between the NNMs (\textcolor{black}{$\boldsymbol{-}$}) and the phase quadrature curves (\textcolor{red}{$\boldsymbol{--}$},\textcolor{dark_green}{$\boldsymbol{--}$}) obtained by applying a single-harmonic excitation at the locations marked in the structure schematic (\textcolor{purple}{$\boldsymbol{\bullet}$},\textcolor{orange}{$\boldsymbol{\bullet}$}). (b) Amplitudes of the physical forces required to isolate the first and second NNMs. Force amplitude curves are coloured according to their respective excitation locations.}
\caption{Forcing amplitudes required to isolate (a) the first and (b) the second NNMs with two input forces. The first excitation considers one force on the small cross beam (\textcolor{purple}{$\boldsymbol{-}$}) and one on the main beam (\textcolor{orange}{$\boldsymbol{-}$}) as in Figure~\ref{fig:flnm12_nnms_freeratio_phys}(a). The second excitation uses forces at both tips of the cross beam (\textcolor{gray}{$\boldsymbol{-}$},\textcolor{ocean}{$\boldsymbol{-}$}). Analytical predictions for the first excitation ($\boldsymbol{\Box}$, $\boldsymbol{\Diamond}$) agree very well with numerical continuation results (solid lines).}
\caption{Modal input force ratios required to obtain the first (\textcolor{pink}{$\hrectangleblack$}) and second (\textcolor{lblue}{$\hrectangleblack$}) NNMs up to a frequency of 20 Hz. Coloured bullet points show the force obtained for the excitation locations reported on the structure schematic. (\textcolor{black}{$\boldsymbol{-}$}) All possible modal force ratios achievable with an excitation applied to only one vertical DOF of the finite element model.}
\caption{Force amplitude required to isolate the first (a) and second (b) NNMs of the clamped-clamped cross beam structure in the case of a single-point, single-harmonic excitation applied to the green location (\textcolor{green}{$\boldsymbol{\bullet}$}). Analytical predictions given by Eq.~\eqref{eq:F1} (\textcolor{black}{$\boldsymbol{--}$}) and results obtained using numerical continuation (\textcolor{green}{$\boldsymbol{-}$}).}
\caption{Regions with large phase errors predict where quadrature curves will be inaccurate. NNMs (solid lines) are coloured according to the magnitude of the phase error. All phase errors larger than $\pi/4$ are coloured in black. Quadrature curves (dashed lines) are coloured according to the excitation location in Figure~\ref{fig:11beam_fphys_circle}: (a) main cross beam (\textcolor{gray}{$\boldsymbol{\bullet}$}), (b) small cross beam (\textcolor{blue}{$\boldsymbol{\bullet}$}). The intersection between the quadrature curve and the second NNM is at 16.69 Hz ($\boldsymbol{+}$).}
\caption{Comparison between the NNMs of the structure (coloured according to the magnitude of the phase error) and the quadrature curves numerically obtained when damping ratios are ten times smaller. Single-point harmonic excitation applied vertically to the orange location (\textcolor{orange}{$\boldsymbol{\bullet}$}). All phase errors larger than $\pi/4$ are coloured in black.}
\caption{Optimization path of a \snet\attached network. The numbers in blue boxes show the order of gradient update.}
\caption{Comparison of the test accuracy of main network ({\color{blue}Blue}) and \snet\augmented network ({\color{red}Red})}
\caption{Comparison with PER ({\color{green}Green}), Baseline (the main network only, {\color{blue}Blue}) and \snet\augmented Network ({\color{red}Red})}
\caption{Learning curves for the cart-pole problem by average reward. Extension of combination of \snet\and PER ({\color{orange}Orange}) and stochastic sampling using \snet\({\color{violet}Purple})}
\caption{Accuracies of different \snet\configurations (amount of parameter shared) on CIFAR10 dataset with small architecture for the main network.}
\caption{Studying a single filter at layer conv5\_fusion: (a) and (b) show what maximizes the unit at the input: multiple coloured blobs in the appearance input (a) and moving circular objects at the motion input (b). (c) shows a sample clip from the test set, and (d) the corresponding optical flow (where the RGB channels correspond to the horizontal, vertical and magnitude flow components respectively). Note that (a) and (b) are optimized from white noise under regularized spatiotemporal variation.\color{cyan}{\bf Best viewed in Adobe Reader where (b)-(d) should play as videos.} }
\caption{Studying the Billiards unit at layer conv5\_fusion from\figref{fig_vis_billiard}. We now show what highly activates the filter in the appearance and in the motion input space using intermediate spatiotemporal variation regularization (a) and (b): Figs.~(c) and (d) show what excites the filter when there is no restriction on the temporal variation of the input: The appearance, (\ref{fig_bil_app_fast}) now also shows a black dot with skin-coloured surroundings at the top which might resemble a head and the motion filter (d) now detects exploding motion patterns (\eg when the white ball hits the others after it has been accelerated by the billiard cue). \color{cyan}{\bf All videos play on click.} }
\caption{Steps in the creation of the \multimarker. The 3-D printed template serves as a housing for the marker and is rigidly attached to a carbon fiber rod such that the marker can be safely introduced into the \xray field of view.}
\caption{Some qualitative results from: NLOR (left column), Ours(I,D,O) (middle) and Ours(I,D,O,G) (right column). These results are obtained on the Cityscapes. {\color{green} Green}: ground truth box and {\color{red}Red}: predicted box.}
\caption{Overall results on Cityscapes. Input descriptions on the first row and input videoframe and gaze in the second row. Middle row represents intermediate results where Gaze estimation is embedded along with object proposals while bottom row represents the final OR results. {\color{green} Green}: ground truth box and {\color{red}Red}: proposals and the predicted boxes.}
\caption{\textcolor{blue}{Example of .... B vs $A - \left(KK^T\right)/|K| $ ......Interpret right figure as propensity to connect?}}
\caption{Sequence tagger F1 using CRF with all features, CCM with all features \& constraints, and semi-supervised CCM over partially labeled crowd data. The second set of results mirror these numbers using a Bi-directional LSTM CRF. Results are statistically significant (paired t-test,p value$<$0.000124). }
\caption{Some sample questions from our test set and the answers returned by our system. %Answer entity names are in the `place\_name' field in the JSON Response. Answers in {\color{cadmiumgreen} green} are identified as correct while those in {\color{red} red} are incorrect.}
\caption{\label{fig:Fig6} Diffusivity as a function of normalised distance from criticality, $D(\epsilon)$, measured by fitting trail intensity profiles ({\color{green} $\blacksquare$}) and from differential dynamic microscopy ({\color{red} $\bullet$}). Lines are fits to $D \sim \epsilon^\phi$ with $\phi = 0.56 \pm 0.08$ and $0.60 \pm 0.03$ respectively.}
\caption{A comparison between our method and the current top methods in CDnet2014 benchmark. Note that these results obtained from CDnet 2014 challenge website\protect\textcolor{magenta}{\textsuperscript{\ref{1}}}.}
\caption{Feature distances calculated and $N$ closest to $c_s$ selected ({\color{red}red}).}
\caption{Given $N$ from Fig. \ref{fig:K1}, $M$ farthest from $c_\ell$ kept ({\color{green}green}).}
\caption{General results for mobility. Upper values are for weekdays and \red{lower ones for weekends} (in red color). \textbf{LJM}: maximum jump [m]; \textbf{DIA}: diameter [m]; \textbf{TJM}: total trajectory length [m]; \textbf{GYR}: radius of gyration [m]; \textbf{BLD}: no. uniq. buildings; \textbf{APC}: access point count; \textbf{PDT}: time spent at preferred building [minutes]; \textbf{DLT}: total session time at each building.}
\caption{Traffic features used for integrated mobility-traffic analysis (per device, per day; see Fig. \ref{fig:correlations} for abbreviations). Upper values are for weekdays and \red{lower ones for weekends} (in red).}
\caption{ Effects of (a) noise level ($\nu$) and (b) $l_1$-penalty parameter ($\lambda_1$) on the estimation error. The solid symbol ({\color{red}$\bullet$}) denotes $(\Delta = 5,P=5)$ and the hollow symbol ($\circ$) is for $(\Delta = 8,P=8)$ }
\caption{\label{fig:8} Example translations where subjective-case pronouns in brackets are dropped in original input but labeled by DP generator. We italicize some {\em \color{blue} mis-translated} errors and highlight the {\bf \color{red} correct} ones in bold.}
\caption{\label{tab1} Comparison of SKR obtainable and time required for 1-decoy and {\color{gray2} 2-decoys} using two different PA block sizes.}
\caption{ Visualization of the space of states of the SIR bivariate Markov Chain $\{s, i\}$. $s, i \in \mathbbm{N}$ and $s + i \leq N $, in other words the space of valid states sits under the green line $s+i = N$. The initial state in an SIR epidemic is always on the \textcolor{ForestGreen}{green line} ($S(0) + I(0) = N$). Given an initial state $\{s = S(0), i = I(0)\}$ (shown by the \textcolor{DarkMagenta}{magenta circle}), the \textcolor{DarkOrange}{orange area} shows the space of reachable states and the \textcolor{gray}{gray area} depicts the \emph{unreachable states}. From a state $\{s,i\}$ the system can to $\{s-1,i+1\}$ %with probability $p_{(s-1,i+1),(s,i)}$ (new infection); and to $\{s,i-1\}$ %with probability $p_{(s,i-1),(s,i)}$ (new recovery). The absorbing states ($\{s, 0\}$) is shown with a \textcolor{red}{red line}. }
\caption{(a) Bar graph of $\xi_\mathrm{u}$ and of its components vs.\the cell efficiency for the three devices. The gap between the bars and the dashed line is the non-thermal lost power fraction.\red{See Appendix A for the $L_\mathrm{4}$ components.} (b) Spectral dependency of $\xi_\mathrm{u}$ for the three cells.}
\caption{Skin-friction coefficient ($C_f$) as a function of non-dimensional parameter $\frac{\nu x}{Ua^2}$ (a), where results for radius based Reynolds number $Re_a=500$, 1000 and 10000 are shown along with solutions of Seban-Bond-Kelly \citep{seban,kelly}({\color{red} $\blacksquare$}) and Glauert--Lighthill \citep{gl}($\Box$). The present result (eq. \ref{zpglbl}) using $\delta^*$ from SBK ({\color{red}$-$}) and GL ($- -$), show identical $C_f$. $C_f$ as a function of $Re_a$ is compared with the result of \cite{cebeci1970} ($\Box$) for long thin cylinder (large $x/a$), where boundary layer thickness reaches asymptotic value \citep{stewartson}(b). The value obtained from the Blasius solution ($- -$) is also shown in (b) for comparison.}
\caption{A \textcolor{red}{\tt meta\_for} loop nest for adding two matrices.}
\caption{A \textcolor{red}{\tt meta\_for} loop nest for adding two matrices.}
\caption{Some retrieval results for the experiments of Sec.~\ref{subsubsection:retrieval}. The first two lines refer to category and pose retrieval, lines 3-4 to category retrieval and lines 5-6 to pose retrieval. Errors are highlighted in \textcolor{red}{red}. For each pair of figures in query part, the left one is original figure, while the right one is the real image corresponding to generated feature. \label{fig:retrieval}}
\caption{Generation results by manipulating captions. The manipulated parts of texts are highlighted in \textbf{bold} characters, where the types of manipulation is indicated by different colors. {\bf \color{blue} Blue}: scene context, {\bf \color{magenta} Magenta}: spatial location, {\bf \color{red} Red}: the number of objects, {\bf \color{green} Green}: object category. }
\caption{Accuracy by train epochs. Curves in \textcolor{blue}{blue} means the train of these two networks without Dropout. Curves in \textcolor{red}{red} denotes the Dropout version of the corresponding models. These accuracies are all calculated from the training data, while the solid curve is under train mode and the dashed one is under evaluation mode. We observe the significant accuracy shift when a network with Dropout ratio $0.5$ changes its state from train to test stage, with all network parameters fixed but the test policies of Dropout and BN applied. }
\caption{Adjust BN's moving mean/variance by running moving average algorithm on training data under test mode. These numbers are all averaged from $5$ parallel runnings with different random initial seeds. \textcolor{red}{\textbf{}}}
\caption{Error rates after applying Dropout after all BN layers. These numbers are all averaged from $5$ parallel runnings with different random initial seeds. \textcolor{red}{\textbf{}}}
\caption{Error rates after applying Dropout after all BN layers on the representative state-of-the-art models on ImageNet. These numbers are averaged from $5$ parallel runnings with different random initial seeds. Consistent improvements can be observed.\textcolor{red}{\textbf{}}}
\caption{Apply new form of Dropout (i.e. Uout) in {Dropout-(b)} models. These numbers are all averaged from $5$ parallel runnings with different random initial seeds. \textcolor{red}{\textbf{}}}
\caption{The backscattering of the incident wave $\ee^{-0.1(x-x_R - t)^2}$, received at $(x_R,0)$, from simulations where there are no particles that took longer than time {\color{red}{40}}, {\color{purple}{100}} and {\color{blue}{180}} for their first scattered wave to arrive at $(x_R,0)$. Figure~\ref{fig:TimeOfFlight} shows the configuration of these particles.}
\caption{Variability of the radio emission obtained from our interferometric observations. Each data point represents single scan during the phase-referencing observations (\textcolor{red}{$\bullet$} represents J1809+5007 and $\mathbin{\vcenter{\hbox{\rule{0.8ex}{0.8ex}}}}$ represents AM\,Her, respectively) . The error bars are of length$\pm1\sigma$.}
\caption{(a) Formation and detection mechanism for ultracold $^{85}$Rb$^{133}$Cs molecules in the X$^{1}$$\Sigma$$^{+}$ (v'' = 0) state. The potential energy curves are based on the data in ref.~{\color{blue}\cite{FahsAlloucheKorekEtAl2002}}. The related pathways are the same with ref.\cite{BruzewiczGustavssonShimasakiEtAl2014}. (b) Brief diagram of optical elements near vacuum chamber. (c) Experimental time sequence. The shadow with different color means ``on'', the blank means ``off'', and the rectangle with lines means ``selectively on or off''. When the lasers are kept on in order to push the co-trapped cold atoms out of the ODT while the lasers are turned off in order to keep the trapped cold atoms in the ODT. }
\caption{Some annotated sequences of the proposed dataset. The target's ground-truth bounding box in the first frame is shown. The sequences are ranked according to their difficulty degree, which is the average TRE score of the tested trackers on each sequence (see the \textbf{supplemental material}). The sequence name, difficulty degree, and corresponding attributes are shown below the image, respectively. {\color{blue}Blue} font denotes old sequences used in the previous studies, while {\color{red} red} font denotes newly collected sequence. The signs, such as, OCC, SV and BC, indicates the attributes of one sequence. More information of these signs can be found in Table \ref{attributes}.}
\caption{Estimated DI graph between the regions: US, Asia, and Europe \textcolor{\changeColor}{with daily sampling}.}
\caption{Cross Domain Transfer: (a) Source Task: Cart-Pole and (b) Target Task:Bicycle, VRML in MATLAB\textregistered is used as simulation environment \cite{bicycle_sim} to demonstrate cross domain transfer.}
\caption{\textcolor{red}{Sample 3D body pose estimation. First box explains the 3D voxel reconstruction segmentation with the segmentation and extracting the voxel data as input of the tracking box. Prediction by using the special filtering, voxel labeling and finally updating by using the filtering approach of the prediction step ~\cite{mikic2003human}.}}
\caption{Prior $\text{p} \left( \kappa_{mk} \vert \tau, \sigma_{\beta_m} \right)$ on $\kappa_mk$ for different values of $\sigma_{\beta_m}$ and $\tau$. The prior distribution skews towards 1 if $\tau$ increases or $\sigma_{\beta_m}$ decreases.} \label{fig1} \end{figure} \begin{figure}[!t] %\vspace{.3in} \centering \includegraphics[scale=0.2]{uniform_vs_beta} \includegraphics[scale=0.27]{pathway_perturbation_distribution} %\vspace{.3in} \caption{Left : Boxplots of Area Under the Curve (AUC) for pathway perturbation inference with uniform prior on $\boldsymbol{\phi}$ compared to a beta like prior on $\boldsymbol{\phi}$ for 10 simulated datasets. We infer perturbations based on the posterior distribution of $ \boldsymbol{\phi}^{\text{controls}} - \boldsymbol{\phi}^{\text{cases}}$. Right : Distribution of $ \boldsymbol{\phi}^{\text{controls}} - \boldsymbol{\phi}^{\text{cases}}$ under the uniform prior (white) and the beta like prior (grey). True perturbed pathways are printed in bold on the x axis.} \label{fig2} \end{figure} \subsection{Integrative analysis} Interactions between heterogeneous omic variables such as transcripts and metabolites or gene expression and metabolites is modeled via the following hierarchical shrinkage model: \begin{eqnarray} \mu_{itm} = \alpha_m + \gamma_{im} + \boldsymbol{y}_{it}\boldsymbol{\beta}_m + \nu_{itm}\\ \label{eq8} \beta_{mk} \vert \lambda_{mk}, \sigma_{\beta_m} \sim N \left( 0, \lambda_{mk}^2 \sigma^2_{\beta_m} \right)\\ \lambda_{mk} \vert \tau \sim \text{St}^+ \left( \tau ,0,1 \right) %\tau \sim \text{C}^+ \left(0,1 \right) \end{eqnarray} where $\text{St}^+$ denotes the half Student-t distribution with $\tau$ degrees of freedom, $\alpha_m$ represents treatment effect for metabolite $m$, $ \gamma_{im} \sim N\left(0, \sigma_{\gamma_m}^2 \right)$ represents individual perturbations for metabolite $m$, $\nu_{itm} \vert \nu_{i,t-1,m} \sim N\left(\theta_m \nu_{i,t-1,m}, \sigma_{\nu_m}^2\right)$ follows and auto-regressive process and represents temporal effects for metabolite $m$ of individual $i$ at time point $t$. Finally, $\boldsymbol{\beta}_m$ quantifies interactions between metabolite $m$ and other omic variables. $\lambda_{mk}$ is called local shrinkage parameter whilst $\sigma^2_{\beta_m}$ is the global shrinkage parameter. For $\tau=1$, this prior reduces to the horseshoe prior. Intuitively, for small values of $\lambda_{mk}$ the coefficient $\beta_{mk}$ is very close to $0$ while for relevant variables $\lambda_{mk}$ will be large. In addition, $\sigma_{\beta_m}$ controls the overall shrinkage level i.e sparsity of the vector $\boldsymbol{\beta}_m$ is more important for small values of $\sigma_{\beta_m}$. Define $\kappa_{mk} = \dfrac{1}{1+ \lambda_{mk}^2 \sigma^2_{\beta_m}/ \tau}$ a random shrinkage coefficient such that $\kappa_{km}\approx 0$ when $\lambda_{mk}$ is large and $\kappa_{km}\approx 1$ when $\lambda_{mk}$ is small. This transformation implies the following prior distribution on $\kappa_{mk}$: \begin{equation} \text{p} \left( \kappa_{mk} \vert \tau, \sigma_{\beta_m} \right) = \dfrac{1}{2 \sqrt{\pi} \textbf{B} \left( \frac{\tau}{2}, \frac{1}{2} \right)} \dfrac{\sigma_{\beta_m}^{\tau} \kappa_{mk}^{\tau/2-1} \left( 1-\kappa_{mk}\right)^{-1/2}}{\left( 1- \kappa_{mk} + \kappa_{mk} \sigma_{\beta_m}^2 \right)} \end{equation} This prior density is shown in figure \ref{fig1} for different values of $\sigma_{\beta_m}$ and $\tau$. It reduces to a Beta$\left(\tau/2, 1/2 \right)$ distribution if $\sigma_{\beta_m}=1$ and to a Beta$\left( 1/2, 1/2 \right)$ which looks like a horseshoe, if in addition $\tau=1$. When $\tau$ increases, Beta$\left(\tau/2, 1/2 \right)$ skews towards $1$ which increases the global shrinkage power. The expectation of $ \boldsymbol{\beta}_m $ given $\boldsymbol{Y}, \boldsymbol{\kappa}_m, \tau, \boldsymbol{\mu}_{tm} $ can be expressed as: \begin{eqnarray} \mathbb{E} \left(\boldsymbol{\beta}_m \vert \boldsymbol{Y}, \boldsymbol{\kappa}_m, \tau, \boldsymbol{\mu}_{tm} \right)&=& \left(\sum_{t=1}^T \boldsymbol{Y}_t^T \Sigma_m^{-1} \boldsymbol{Y}_t + \tau \Upsilon_m \right)^{-1} \nonumber \\ & & \times \sum_{t=1}^T \boldsymbol{Y}_t^T \Sigma_m^{-1} \boldsymbol{\mu}_{tm} \label{eq12} \end{eqnarray} where $\Sigma_m = \left( \frac{\sigma_{\nu_m}^2}{1-\theta_m^2} + \sigma_{\gamma_m}^2 \right) \boldsymbol{I}_N$ and $\Upsilon_m$ is a diagonal matrix of order $K$ with elements $1/\kappa_{mk}-1$. Equation~(\ref{eq12}) introduces a penalty term $ \tau \Upsilon_m$ where $\Upsilon_m$ is a metabolite specific penalty term introduced by the horseshoe prior and $\tau$ is a global penalty term. Precisely, $\tau$ captures the overall sparsity level amongst all metabolites. The expectation of $ \boldsymbol{\beta}_m $ given $\boldsymbol{Y}, \boldsymbol{\kappa}_m, \tau, \boldsymbol{\mu}_{tm} $ is very similar to the estimate of $ \boldsymbol{\beta}_m $ under ridge regression where $ \tau \Upsilon_m$ simply reduces to $\tau \boldsymbol{I}_N$.\\ The global sparsity level can be controlled using $\tau$. Increasing the global sparsity level is a desired property in omic studies, as usually we deal with a large number of omic variables where only few are important. Moreover, when there is prior knowledge available, specifying $\tau$ \textit{a priori} can optimize the inference and additionally, provide a more informative prior on $\lambda_{mk}$. If we fix $\text{p}\left( \sigma_{\beta_m}^2\right) \propto 1/\sigma_{\beta_m}^2$, integrating over $\sigma_{\beta_m}$ gives the expected value of $\kappa_{mk}$ as : \begin{eqnarray*} \mathbb{E} \left(\kappa_{km} \vert \tau \right)= \frac{\Gamma \left(1/2\right)^{-1}}{2 \sqrt{\pi} \Gamma \left(\tau/2\right)} \\ \times \textbf{G}_{3,3}^{2,3} \left(\begin{matrix} 1, \tau/2, 0 \hfill \\ \tau/2, \tau/2-1/2, 0 \end{matrix} \,\middle\vert \, 1-\sigma_{\beta_m}^2 \right) \label{eq9} \end{eqnarray*} where $\textbf{G}_{\cdot,\cdot}^{\cdot,\cdot} $ is Meijer's G-function \citep{meijer1936}. The equation above %~(\ref{eq9}) can be used to fix $\tau$ \textit{a priori} by defining the expected proportion of shrunk coefficients. In practice, different values of $\tau$ are plugged into the equation above %~(\ref{eq9}) to get the desired proportion of shrunk coefficients. However, many definite integrals can be obtained using the tables of Meijer functions in \cite{brychkov2008} for special values of parameters. \begin{figure}[!t] %\vspace{.3in} \centering \includegraphics[scale=0.3]{pathway_perturbation_roc} %\vspace{.3in} \caption{Average Receiver Operating Characteristic (ROC) curves for pathway perturbation inference across 10 datasets for different factions of ``falsely" assigned metabolites.} \label{fig5} \end{figure} %\begin{figure}[!t] %%\vspace{.3in} %\centering %\includegraphics[scale=0.3]{pathway_perturbation_boxplot} %%\vspace{.3in} %\caption{Right : Boxplots of Area Under the Curve (AUC) for pathway perturbation inference for different fractions of perturbed metabolites. Perturbed metabolites are falsely assigned to no pathway or different pathways. Our model performs significantly well (above 0.7 on average) if less than $44\%$ of metabolites in each pathway are perturbed and reduces to random guess if $50\%$ of metabolites are perturbed.} %\label{fig6} %\end{figure} \subsection{Experimental design} The covariance structure between metabolites might change drastically as a result of treatment if the latter affects metabolic pathways. The model can be extended to take into account the experimental design. As specified in the previous section, $\alpha_m$ captures the treatment effect for metabolite $m$, $ \gamma_{im}$ represents individual perturbations for metabolite $m$, $\nu_{itm} \vert \nu_{i,t-1,m} \sim N\left(\theta_m \nu_{i,t-1,m}, \sigma_{\nu_m}^2\right)$ represents temporal effects for metabolite $m$ of individual $i$ at time point $t$ in equation~(\ref{eq8}). In addition, we allow covariance structures $\boldsymbol{C} \left(\boldsymbol{\phi}^e \right)$ to be different for the control samples and the cases where $e \in \lbrace \text{cases, controls} \rbrace$ designates experimental groups. This yields the overall hierarchical model: \begin{eqnarray} \boldsymbol{x}_{it}^e \vert \boldsymbol{\mu}_{it}, \boldsymbol{C}, \sigma \sim N \left(\boldsymbol{\mu}_{it} , \left(\boldsymbol{I}_M-\boldsymbol{C \left(\boldsymbol{\phi}^e\right)} \right)^{-1} \sigma^2 \right)\label{eqx}\\ \mu_{itm} = \alpha_m + \gamma_{im} + \boldsymbol{y}_{it}\boldsymbol{\beta}_m + \nu_{itm} \label{eqmu}\\ \beta_{mk} \vert \lambda_{mk}, \sigma_{\beta_m} \sim N \left(0, \lambda_{mk}^2 \sigma^2_{\beta_m} \right)\\\lambda_{mk} \vert \tau \sim \text{St}^+ \left(\tau ,0,1 \right)\\\gamma_{im} \vert \sigma_{\gamma_m} \sim N\left(0, \sigma_{\gamma_m}^2 \right)\label{eqgam}\\ \nu_{itm} \vert \theta_m, \sigma_{\nu_m} \sim N\left(\theta_m \nu_{i,t-1,m}, \sigma_{\nu_m}^2\right)\label{eqnu} \end{eqnarray} Note that by specifying different dependence parameters for metabolite interactions in cases and controls metabolism, the model is able to identify perturbed pathways by comparing $ \boldsymbol{\phi}^{\text{cases}} $ and $\boldsymbol{\phi}^{\text{controls}}$. %\vspace*{-10pt} \section{RESULTS} In this section we perform experiments on both synthetic and real data to investigate whether our algorithm gives reasonable solutions. We first try our method on a simulated dataset in section \ref{secsynt} to get an understanding of the performance of our method. In section \ref{secmet}, we test our method on a data set using metabolomic and bacterial composition in a drug treatment experiment. In the following, we refer to our model as `` iCARH " model for `` integrative CAR Horseshoe " model. \subsection{Simulation study} \label{secsynt} To get better understanding of our method and test its applicability, we first perform our approach on synthetic datasets. We will mainly focus on the ability of our model to infer pathway perturbation. \subsubsection*{Assessing pathway inference with beta like prior} In the first simulation our objective is to assess how the beta like prior in equation (\ref{eqphip}) improves the iCARH model. We first fixed the number of pathways P to 11, then simulate the design matrices $\boldsymbol{A_p}$. Specifically, a membership matrix $\boldsymbol{Z}$ with dimensions $M \times P$ is randomly generated based on the density of the number of KEGG pathways in which a single metabolite is involved. Each design matrix $\boldsymbol{A_p}$ is then equal to $\boldsymbol{z_p}\boldsymbol{z_p}^T$ where $\boldsymbol{z_p}$ is the $p$th column of $Z$. Finally, we generated 10 datasets according to the model below in order to assess how our model infers perturbed pathways: \begin{eqnarray} \omega \sim \text{Bernoulli} \left(\pi_{\omega} \right)\\\resizebox{0.95 \linewidth}{!}{ $\phi_p^{\text{controls}} \vert \omega \sim \omega N_{[0,1/P\xi_p^2]} \left(\frac{1}{P\xi_p^2}-\rho , \sigma_{\phi}^2 \right) + \left( 1-\omega \right) N \left(0,\psi^2 \right)$ }\\ \resizebox{.95 \linewidth}{!} { $\phi_p^{\text{cases}} \vert \phi_p^{\text{controls}}, \omega \sim \omega N_{[1/P\xi_p^1,0]} \left(\frac{1}{P\xi_p^1}+\rho , \sigma_{\phi}^2 \right)+ \left(1-\omega \right)\delta_{\phi_p^\text{controls}} $ } \end{eqnarray} with $N_{[a,b]}$ denotes the truncated normal distribution with boundaries $a$, $b$, $\xi_p^1$ and $\xi_p^2$ are the minimum and maximum eigenvalues of $\boldsymbol{G_p}\boldsymbol{A_p}$. The rest of the parameters is set as follows : number of bacterial variables $K=1$, number of metabolites $M=40$, number of time points $T=7$, number of samples $N=22$, global parameter $\tau$ fixed to~1.2, parameters $\nu_{itm}$, $\gamma_{im}$, $ \mu_{itm}$, $ \boldsymbol{x}_{it}^e$ simulated according to equations~(\ref{eqnu}),~(\ref{eqgam}),~(\ref{eqmu}),~(\ref{eqx}) respectively. %\begin{figure}[!t] %%\vspace{.3in} %\centering %\includegraphics[scale=0.5]{weighted_error_pathways.jpg} %%\vspace{.3in} %\caption{Weighted error of the model under uniform distribution on $\boldsymbol{\phi}$ (blue) and beta distribution (red). The x-axis represents various threshold settings of the true positive rate (TPR). The weighted error is obtained using the weighted false positive rate (FPR) and weighted false negative rate (FNR). Inferring perturbed pathways is slightly improved by using the beta distribution.} %\label{fig2} %\end{figure} \begin{figure}[h] %\vspace{.3in} \centering \includegraphics[scale=0.4]{vsdppca}%height=6cm, width=8cm %\vspace{.3in} \caption{Posterior predictive checks for mean absolute deviation (MAD) compared to DPPCA for different numbers of metabolites included. The MAD decreases as the number of metabolites increases. Our model performs clearly better than the DPPCA model.} \label{fig3} \end{figure} We set non-informative uniform priors on $\alpha_m$, $\sigma_{\gamma_{im}}$, $\theta_m$, $\sigma_{\mu_m}^2$. We set an informative prior on $\sigma^2_{\gamma_m} \sim \text{inverse-gamma} \left(1, 0.1 \right)$ as we expect low variability amongst biological samples of the same group. We fix $\pi_{\omega}$ to $0.7$ the proportion of expected perturbed pathways. $\rho$ is fixed to a value of $0.05$, $\sigma_{\phi}^2$ to $0.2$ and $\psi$ to a large value. We compare inference of the model under a uniform prior for $\phi^e_p$ and the prior in equation ~(\ref{eqphip}). Inference is done using 2000 iterations of Hamiltonian Monte Carlo sampling and 1000 warm-up iterations. %\begin{figure}[!h] %\vspace{.3in} %\includegraphics[height=9cm, width=8cm]{car_vs_sme_202.jpg} %\vspace{.3in} %\caption{Absolute mean error of estimated metabolic profiles using our model (red) and the SME model (green). Our model performs in a similar way to SME which is indicative of a good fit.} %\label{fig3} %\end{figure} The left plot in figure \ref{fig2} shows the boxplots of the Area Under the Curve (AUC) for pathway perturbation inference for 10 simulated datasets with uniform prior on $\boldsymbol{\phi}$ and a beta like prior on $\boldsymbol{\phi}$. We infer perturbations based on the posterior probability that $ \boldsymbol{\phi}^{\text{controls}}$ and $\boldsymbol{\phi}^{\text{cases}}$ are different i.e. the $95\%$ credible interval of $ \boldsymbol{\phi}^{\text{controls}} - \boldsymbol{\phi}^{\text{cases}}$ does not contain zero. The AUC values for the beta like distribution is significantly higher than the AUC values for the uniform distribution. On average pathway peturbation inference under the uniform distribution reduces to a random guess with an average AUC of 0.53. This is likely due to the lack of variance of the uniform distribution. The right plot in figure \ref{fig2} shows the posterior distributions of $ \boldsymbol{\phi}^{\text{cases}} - \boldsymbol{\phi}^{\text{cases}}$ under the uniform prior (white) and the beta like prior (grey) for each pathway. True perturbed pathways are printed in bold on the x axis. \subsubsection*{Assessing pathway inference against design inaccuracies} It is very common in metabolomics data to find metabolites that are correlated but not in the same KEGG pathway. In the following simulation we assess how inaccuracies in the covariance structure between metabolites and the design matrices $\boldsymbol{A_p}$ affect the iCARH model. We used the 10 datasets from the previous simulation and perturbed the design matrices by selecting a random fraction of metabolites in each pathway. We then randomly (falsely) assign these metabolites to no pathway, or to different pathways. We similarly run the model for 2000 iterations of Hamiltonian Monte Carlo sampling and 1000 warm-up iterations for each of the fractions $y = 0, 0.18, 0.35, 0.44, 0.5, 0.62$ of perturbed metabolites. Finally, in the same fashion, we assess perturbations based on the $95\%$ credible interval of $ \boldsymbol{\phi}^{\text{controls}} - \boldsymbol{\phi}^{\text{cases}}$. Figure \ref{fig5} is a series of average Receiver Operating Characteristic (ROC) curves across 10 datasets for each of the factions $y$. On average, the performance of our model reduces to a random guess (AUC of 0.5) if $50\%$ of the metabolites in each pathway is perturbed. The AUC of our model reaches 0.97 if no metabolites are perturbed and is about 0.88 if $18\%$ of the metabolites in each pathway are perturbed. \begin{figure}[!t] %\vspace{.3in} \centering \includegraphics[scale=0.3]{qqplot}%height=6cm, width=8cm %\vspace{.3in} \caption{Right and left panels show model fit assessment for controls and cases data. Left : quantile-quantile normal plot of $\Psi^{-1}_{\text{cases}} \left( \boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right)$. Right : quantile-quantile normal plot of $\Psi^{-1}_{\text{controls}} \left(\boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right)$.} \label{fig7} \end{figure} \subsection{Case study} \label{secmet} In this section, we test our model on an actual metabolomic data and 16S data for bacterial profiles. In this study we are interested in the influence of metformin on a non-diabetic model. Metformin is the first-line medicine to treat type 2 diabetes. It has also been suggested that metformin has anti-cancer, cardiovascular and anti-aging effects. Because of their very large metabolic capacity, the gut bacteria can influence toxicity and metabolism of drugs. Here, we are particularly looking for metabolic biomarkers indicative of microbiota changes as result of treatment. The study design is as follows: metabolic profiles of 24 rats are acquired once a week using different mass spectrometry techniques from plasma samples over a period of 9 weeks. Bacterial profiles are acquired using miSeq. The study has allowed for two groups of 12 rats where metformin has been administrated to the second group (weeks 3 to 7) allowing for 2 weeks of acclimatation (weeks 1 and 2) and 2 weeks of recovery (weeks 8 and 9). After data processing and metabolite identification, a total of 56 metabolites and 6 bacteria species are further analysed using our model. Inference is done using 2000 iterations of Hamiltonian Monte Carlo sampling. We assess performance of our model for different values of $\tau$ using the Watanabe-Akaike information criterion (WAIC). Tested values of $\tau$ comprise 1, 1.2, 5, 10 with corresponding WAIC values of 7317.296 , 7322.798 , 7317.457 , 7316.476 respectively. WAIC values are very similar for different values of $\tau$ which suggests to use the most selective model with $\tau=10$ as it is the simplest i.e with the smallest number of selected variables. %\begin{table}[!h] %\processtable{WAIC for different values of $\tau$\label{tab1}} {\begin{tabular}{@{}lllll@{}}\toprule %$\boldsymbol{\tau}$ & {\bf 1} &{\bf 1.2} &{\bf 5} &{\bf 10} \\ %\midrule %\textbf{WAIC} & 7317.296 & 7322.798 & 7317.457 & 7316.476 \\ %\botrule %\end{tabular}}{}%{The model does not seem to be very sensitive to different input values for $\tau$ in this case. } %\end{table} \begin{figure}[!t] %\vspace{.3in} %\hspace{-0.5cm} \includegraphics[scale=0.4]{treatment_effect_partiel} %\vspace{-.5in} \caption{Estimates of effects of treatment on metabolite profiles are captured by $\beta^{\alpha}_m$. Only part of the data is plotted as we are only interested in ``metabolite 27". } \label{fig9} \end{figure} %We then compare predictive ability of our model with the spline mixed effects (SME) model. The SME model in one of the most popular algorithms in the metabolomics community for analysing time course data. It was designed in order to prevent under- and overfitting by setting the right smoothing penalty. However, fitting is done in a univariate way. We perform cross-validation for both models by predicting metabolites each time using the previous time points. We compare performance of each model using the absolute mean error over predicted time points. % %Figure \ref{fig3} shows the absolute mean error of our model and the SME model. Both models perform in a very similar way. This is indicative of a good fit of our model. \subsubsection*{Assessing model fit} In order to assess our model fit, we perform posterior predictive checks of our model compared to DPPCA \citep{2014dynamic}. The DPPCA model is a multivariate model using PCA, where PCA scores are modeled via a stochastic volatility model. In the Bayesian framework, posterior predictive checks consist in comparing data simulated from the posterior predictive distribution with the observed data. We compared the simulated data and the observed data by means of mean absolute deviations (MADs) between the observed and the simulated covariance matrices for different numbers of metabolites included i.e only part of the data corresponding to these metabolites is considered. The same process was repeated for inference using the DPPCA model \citep{2014dynamic}. Figure \ref{fig3} shows MADs of our model and the DPPCA model. As expected, MADs for both models decrease when the number of metabolites increases. Overall, our model clearly outperforms the DPPCA model. In addition to posterior predictive checks, goodness of fit was also checked by using $\Psi^{-1}_e \left( \boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right) \sim N\left(0, \boldsymbol{I}_M\right)$ where $\Psi_e$ denotes the Cholesky factor of $\left(\boldsymbol{I}_M-\boldsymbol{C \left(\boldsymbol{\phi}^e\right)} \right)^{-1} \sigma^2$. Zero-mean and normality were thus checked for $\Psi^{-1}_e \left(\boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right)$ (See figure \ref{fig7}). \subsubsection*{Data results} Since the administrated drug was also profiled using mass spectrometry, we fit the iCARH model with $\boldsymbol{\alpha}_m = \beta^{\alpha}_m \boldsymbol{y}_{\text{drug}}$. Figure \ref{fig9} is a series of boxplots for $\beta^{\alpha}_m$ for metabolites 13 to 31. We are mainly interested in ``metabolite 27" as it is associated with some bacteria species. Figures \ref{fig8} and \ref{fig4} show posterior distributions of $\boldsymbol{\phi}^e$ for each pathway and estimates of effects of bacteria on metabolites. Results in section \ref{secsynt} suggest to compare the covariance structure of metabolites in the observed data with the covariance induced by the design matrices in order to have an \textit{a priori} idea on the robustness of pathway inference (See figure \ref{fig5}). For a correlation threshold of $0.3$, about $25\%$ of the metabolites are misspecified in the design matrices which corresponds to an AUC around $0.8$ according to figure \ref{fig5}. If we set a higher correlation threshold, a lower number of metabolites will be misspecified. This supports the use of the iCARH model for pathway perturbation inference for this data. Estimates of effects of bacteria on metabolite profiles are captured by $\boldsymbol{\beta}_m$. Some metabolites present significant changes along with the bacterial profiles. For example, ``metabolite 27", a hydroxy fatty acid, is associated with alterations in abundance of 4 bacteria species. Figure \ref{fig8} shows that, as a result of treatment, KEGG pathways are not significantly altered. However, distributions of $ \boldsymbol{\phi}^{\text{controls}}$ for ``fatty acids biosynthesis" and ``biosynthesis of unsaturated fatty acids" KEGG pathways are remarkably flatter than the distributions of $ \boldsymbol{\phi}^{\text{cases}}$. These pathways involve the previously identified hydroxy fatty acid metabolite. Our analysis confirms previously reported studies that hydroxy fatty acids might be produced by the gut microbiome \citep{kishino2013, kimura2013}. \begin{figure}[b] %\vspace{.3in} \centering \includegraphics[scale=0.25]{data_pathways}%height=6cm, width=8cm %\vspace{.3in} \caption{Posterior distributions of $\boldsymbol{\phi}^e$ for each pathway for each treatment group. Posterior distributions of $\boldsymbol{\phi}^e$ are very similar for both controls and cases which is indicative of no -or mild- pathway alterations.} \label{fig8} \end{figure} \section{DISCUSSION} Identifying biomarkers in time course metabolic data and inferring significant associations with heterogeneous omic variables is extremely challenging due to the several sources of variations of the data. In addition, existing methods developed to analyze such data are very scarce and have the limitations of i) overfitting to the few available data points or ii) confounding the experimental and longitudinal variation or iii) ignoring the metabolite interactions or iv) ignoring effects of other omic variables. In this paper, the model we have developed combines several approaches to take into account the different aspects of the data namely the number of time points, the experimental variation captured by $\boldsymbol{\mu}_{it}$, interactions between metabolites captured by $\boldsymbol{\phi}$ and interactions with additional omic variables captured by $\boldsymbol{\beta}_m$. Our results demonstrate that our model successfully addresses the main questions of a metabolomic study. Most importantly, our model is able to identify metabolic biomarkers related to treatment, infer perturbed pathways as a result of treatment and find significant associations with additional omic variables. We have shown that providing an informative prior on metabolic pathways and an informative prior over the parameter $\boldsymbol{\phi}$ is a significant improvement over the DPPCA model. Particularly, our model is more robust to slight variations usually observed in short time series data thanks to the small number of covariance parameters needed to estimate compared to DPPCA. We have also shown through simulation that an informative beta like prior compares better than a non-informative uniform prior in inferring significant pathways. On real data, we have investigated how the number of profiled metabolites can affect the predictive ability of the model. %Using all available measurements, we have also shown that our model yields a remarkably similar predictive performance to SME. Several potential extensions arise naturally from our model. In terms of the metabolite interactions component, many research questions can arise. Alternative strategies to modeling metabolite interactions can be examined such as modeling the non-zero elements of the adjacency matrix $\boldsymbol{C}$ of each pathway as random variables. This strategy was adopted in the CAR literature by \cite{lee2013, Rushworth2017} to take into account step changes in spatial variation. Step changes can potentially be useful to model changes in metabolites correlations as a result of treatment. \cite{lee2011} provide an overview of different CAR models used in spatial modeling. The proposed models can be adapted to fit into the metabolomics literature. From a practical point of view, the model has been fitted using HMC sampling but takes a large amount of time (about 1 hour) mostly because of the variable selection procedure and metabolites interdependence. This could be addressed by using variational Bayes. In fact, variational Bayes inference procedures offer cost-effective inference by means of principled approximations and appealing computational time for high dimensional data. A variational bayes inference of CAR models was proposed by \cite{harrison2010} for high dimensional data, and a variational bayes approach for variable selection was recently proposed by \cite{ormerod2017}. \newpage \begin{figure}[!t] %\vspace{.3in} %\hspace{-0.5cm} \includegraphics[scale=0.3]{bacteria_effect} %\vspace{-.5in} \caption{Estimates of effects of bacteria on metabolite profiles are captured by $\boldsymbol{\beta}_m$.} \label{fig4} \end{figure} \newpage\null\thispagestyle{empty}\newpage \newpage\null\thispagestyle{empty}\newpage \section{CONCLUSION} Metabolomics longitudinal profiling techniques are imperative to understand the effect of a drug or a disease across time and can provide enhanced understanding of the underlying biology of the system. In a data integration framework, we have illustrated the use of the CAR model to incorporate metabolites interactions in the model and the horseshoe prior to identify association with heterogeneous omic variables obtained by other omic techniques. The combination of the CAR and horseshoe levels yields the ``integrative CAR Horseshoe" (iCARH) model which we presented in this article.\\ Although, it is computationally expensive, the iCARH model has various appealing features such that it is able to identify metabolic biomarkers related to treatment, infer perturbed pathways as a result of treatment and identify potential associations between heterogeneous omic variables. Clearly, these appealing features open up further research topics. \subsubsection*{Acknowledgements} Thanks are due to Panagiotis Vorkas for providing the metabolic and 16S data. Infrastructure support for this work was provided by the NIHR Imperial Biomedical Research Centre. \subsubsection*{References} %\bibliographystyle{spmpsci} \printbibliography % \end{document} }}\end{eqnarray}}\end{eqnarray}}\end{eqnarray*}}
\caption{Right and left panels show model fit assessment for controls and cases data. Left : quantile-quantile normal plot of $\Psi^{-1}_{\text{cases}} \left( \boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right)$. Right : quantile-quantile normal plot of $\Psi^{-1}_{\text{controls}} \left(\boldsymbol{x}_{it} - \boldsymbol{\mu}_{it} \right)$.}
\caption{\textcolor{red}{Updated} Ladder CNN with different structures in PAMAP2 dataset. 1-0-0-0-0-0-0-0-0 means we enforce the bottom layer to be similar}
\caption{ActiTrack Dataset: \textcolor{red}{ActiTracker: 100, 500, 5000, 50000}}
\caption{\textcolor{red}{PAMAP2: 100, 500, 1000, 10000}}
\caption{\textcolor{red}{mHealth: 100, 500, 1000, 10000}}
\caption{\footnotesize \textbf{Q}: How many objects are preferred by less than 7 people in at least one category?\\ \textcolor{red}{SAN}: two \cmark ~ \textcolor{blue}{MOM}: two \cmark ~ \textcolor{magenta}{SANDY}: two\cmark\\ \textbf{Q}: What category does the medium purple color represent?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: lisit \xmark ~ \textcolor{magenta}{SANDY}: list \cmark\\ }
\caption{ \footnotesize \textbf{Q}: Which item sold the most number of units summed across all the stores?\\ \textcolor{red}{SAN}: closet\xmark~ \textcolor{blue}{MOM}: branch\cmark~ \textcolor{magenta}{SANDY}: branch\cmark\\ \textbf{Q}: How many units of the item branch were sold in the store sister?\\ \textcolor{red}{SAN}: 9 \cmark ~ \textcolor{blue}{MOM}: 9 \cmark ~ \textcolor{magenta}{SANDY}: 9 \cmark\\ }
\caption{\footnotesize \textbf{Q}: Are the values in the chart presented in a percentage scale?\\ \textcolor{red}{SAN}: no \cmark~ \textcolor{blue}{MOM}: no \cmark~\textcolor{magenta}{SANDY}: no \cmark\\ \textbf{Q}: How many units of items lead and pure were sold?\\ \textcolor{red}{SAN}: 8 \xmark ~ \textcolor{blue}{MOM}: 8 \xmark ~ \textcolor{magenta}{SANDY}: 7 \cmark\\ }
\caption{\textbf{Q}: What is the label of the second bar from the left in each group?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: guest \cmark ~ \textcolor{magenta}{SANDY}: guest \cmark\\ \textbf{Q}: Is each bar a single solid color without patterns?\\ \textcolor{red}{SAN}: yes \cmark~ \textcolor{blue}{MOM}: yes \cmark~ \textcolor{magenta}{SANDY}: yes \cmark \\}
\caption{\textbf{Q}: How many items sold less than 6 units in at least one store?\\ \textcolor{red}{SAN}: four \cmark ~ \textcolor{blue}{MOM}: four \cmark~ \textcolor{magenta}{SANDY}: four\cmark \\ \textbf{Q}: Does the chart contain stacked bars?\\ \textcolor{red}{SAN}: yes \cmark ~ \textcolor{blue}{MOM}: yes \cmark ~ \textcolor{magenta}{SANDY}: yes \cmark\\ }
\caption{\textbf{Q}: What is the highest accuracy reported in the whole chart?\\ \textcolor{red}{SAN}: 7 \cmark ~ \textcolor{blue}{MOM}: 7 \cmark~ \textcolor{magenta}{SANDY}: 7 \cmark\\ \textbf{Q}: Is the accuracy of the algorithm leg in the dataset suite smaller than the accuracy of the algorithm chest in the dataset sample?\\ \textcolor{red}{SAN}: no \xmark ~ \textcolor{blue}{MOM}: no \xmark ~ \textcolor{magenta}{SANDY}: yes \cmark \\ \label{fig:correct-c} }
\caption{\textbf{Q}: Which bar has the largest value?\\ \textcolor{red}{SAN}: closet \xmark ~\textcolor{blue}{MOM}: aspect \cmark ~ \textcolor{magenta}{SANDY}: aspect \cmark\\ \textbf{Q}: What is the value of the largest bar?\\ \textcolor{red}{SAN}: $10^9$ \cmark ~ \textcolor{blue}{MOM}: $10^9$ \cmark ~ \textcolor{magenta}{SANDY}: $10^9$\cmark \\ }
\caption{\textbf{Q}: How many algorithms have accuracy lower than 3 in at least one dataset?\\ \textcolor{red}{SAN}: zero \cmark ~ \textcolor{blue}{MOM}: zero \cmark ~ \textcolor{magenta}{SANDY}: zero \cmark\\ \textbf{Q}: Which algorithm has highest accuracy for any dataset?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: girl \cmark ~ \textcolor{magenta}{SANDY}: girl\cmark}
\caption{\textbf{Q}: Which object is preferred by the most number of people summed across all the categories?\\ \textcolor{red}{SAN}: closet \xmark~ \textcolor{blue}{MOM}: site \cmark~ \textcolor{magenta}{SANDY}: site\cmark\\ \textbf{Q}: Are the bars horizontal?\\ \textcolor{red}{SAN}: yes \cmark ~ \textcolor{blue}{MOM}: yes \cmark ~ \textcolor{magenta}{SANDY}: yes \cmark \\ }
\caption{ \textbf{Q}: What is the label of the third bar from the bottom? \\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: whidkw \xmark ~ \textcolor{magenta}{SANDY}: widow \cmark\\ \label{subfig:a}}
\caption{ \textbf{Q}: Which algorithm has the largest accuracy summed across all the datasets?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: lack \xmark ~ \textcolor{magenta}{SANDY}: vector \cmark\\ \label{subfig:b} }
\caption{\textbf{Q}: Is the value of output smaller than demand? \\ \textcolor{red}{SAN}: no \xmark ~ \textcolor{blue}{MOM}: no \xmark ~ \textcolor{magenta}{SANDY}: yes \cmark\\ \label{subfig:c} }
\caption{\textbf{Q}: Which algorithm has the smallest accuracy summed across all the datasets?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: fil \xmark ~ \textcolor{magenta}{SANDY}: editor \xmark\\ \textbf{Q}: What is the highest accuracy reported in the whole chart? \\ \textcolor{red}{SAN}: 60 \xmark~ \textcolor{blue}{MOM}: 60 \xmark~ \textcolor{magenta}{SANDY}: 60 \xmark\\ \label{subfig:d} }
\caption{\textbf{Q}: How many total people preferred the object terror across all the categories?\\ \textcolor{red}{SAN}: 10 \xmark ~ \textcolor{blue}{MOM}: 10 \xmark ~ \textcolor{magenta}{SANDY}: 10 \xmark\\ \textbf{Q}: How many people prefer the object terror in the category roll?\\ \textcolor{red}{SAN}: 1 \xmark ~ \textcolor{blue}{MOM}: 1 \xmark ~ \textcolor{magenta}{SANDY}: 9 \cmark \\ \label{subfig:e} }
\caption{\textbf{Q}: What is the highest accuracy reported in the whole chart?\\ \textcolor{red}{SAN}: 90 \cmark ~ \textcolor{blue}{MOM}: 90 \cmark ~ \textcolor{magenta}{SANDY}: 80 \xmark \\ \textbf{Q}: Which algorithm has the smallest accuracy summed across all the datasets?\\ \textcolor{red}{SAN}: closet \xmark ~ \textcolor{blue}{MOM}: record \cmark ~ \textcolor{magenta}{SANDY}: park \xmark \\ \label{subfig:f} }
\caption{ Plot of $k$-instanton contribution for $p=3, N=80$ at $\th=\frac{\pi}{3}$. In \subref{sfig:zinst} we show the sum of instanton contributions up to $k$-instantons normalized by $Z_{\text{inst}}$, and in \subref{sfig:zinst-zk} we show each $k$-instanton contribution individually. In both figures \subref{sfig:zinst} and \subref{sfig:zinst-zk}, we used the same colors: $k=1$ ({\color{red}red}), $k=2$ ({\color{green}green}), $k=3$ ({\color{orange}orange}), $k=4$ ({\color{gray}gray}). }
\caption{\label{table9}Connective tissue Raman bands assignment based in $C_{2}$ model. Additional vibrations compared to usual assignment in the fingerprint region are shown in \textcolor{red}{red}. It is important to notice that torsion and stretching at least one of the rings, C-C, CH, CN, NH, C=O, CH$_{3}$, OH stretching, CH$_{2}$ symmetrical and asymmetrical stretching, Amide III, and H$_{2}$O scissoring vibrations appeared in all bands in some degree.}
\caption{Algorithms used and hyperparameters tuned for each. Each (algorithm, hyperparameter) set was used with each of the four feature types, yielding a total of $24 \times 4 = 96$ models. Preprocessing codes: \textasteriskcentered indicates zero-variance predictors, if any, were dropped for model training (as a requirement of model-fitting algorithms); \textdagger indicates predictors were centered and scaled.}
\caption{The master oscillator is an extended cavity diode laser (ECDL) at \SI{972}{\nm} and amplified as in \cite{burkley:2016}. An EOM within the cavity applies a \SI{2.5}{\MHz} dither for locking the three subsequent enhancement cavities. The IR light is frequency quadrupled with two bow-tie enhancement cavities -- first from \ir to \blue using LBO with noncritical (temperature) phase matching, then from \blue to \uv using CLBO with critical phase matching. We characterize the \uv output with a fluorescent screen and CCD camera.}
\caption{\textbf{Methods to Generate Training Label from RECIST:} pixel-wise precision, recall, and mean DICE (mDICE) are reported with standard deviation ($\pm$stdv.). Six different setup configurations are compared, 1) RECIST: dilated RECIST, 2) DCRF: dense CRF, 3) GrabCut: uses only RECIST bbox, 4) GrabCut$^i$: uses b-box and interior foreground, 5) GrabCut-R: uses b-box and dilated RECIST, 6) GrabCut-R-SR: implements GrabCut-R on the super-resolution images. See Sec.~\textcolor{red}{3.1} and experiment settings for details.}
\caption{Simulation results for the type-I error level for $\varphi^\mu$ $(\textendash\textendash\textendash)$, $T_{ML}$ ({\color{black}$--$}), and $T_{tw}$ ({\color{ao(english)}$-\cdot-$}) for different distributions under varying correlation factors $\vSigma=\vSigma_1$ (left) and $\vSigma=\vSigma_1$ (right) for testing $H_0^\mu:\{\mu_1=\mu_2\}$ with $n=10$ (left) and $n=20$ (right) and different missing percentages $r\in\{10\%, 30\%\}$ under the MCAR framework.}
\caption{Simulation results for the type-I error level for $\varphi^{(p)}$ $(\textendash\textendash\textendash)$, $T_{KHLB}(t-App)$ ({\color{black}$--$}), and $T_{KHLB}(normal)$ ({\color{ao(english)}$-\cdot-$}) for different distributions under varying correlation factors {\color{black}($\vSigma=\vSigma_1$)} for testing $H_0^p:\{p=1/2\}$ with $n=10$ (left) and $n=20$ (right) and different missing percentages $r\in\{10\%, 30\%\}$ under the MCAR framework.}
\caption{Simulation results for the type-I error level for $\varphi^{(p)}$ $(\textendash\textendash\textendash)$, $T_{KHLB}(t-App)$ ({\color{blue}$--$}), and $T_{KHLB}(normal)$ ({\color{ao(english)}$-\cdot-$}) for different distributions under varying correlation factors {\color{black}($\vSigma=\vSigma_2$)} for testing $H_0^p:\{p=1/2\}$ with $n=10$ (left) and $n=20$ (right) and different missing percentages $r\in\{10\%, 30\%\}$ under the MCAR framework.}
\caption{Power simulation results $(\alpha=0.05)$ for $T^P$ $(\textendash\textendash\textendash)$, $T_{KHLB}(t-app)$ ({\color{blue}$--$}), and $T_{KHLB}(normal)$ ({\color{ao(english)}$-\cdot-$}) for Symmetric and Asymmetric distributions under $\rho=0.5$ for testing $H_0^p:\{p=1/2\}$ with $n=10$ (left) and $n=20$ (right) and different percentages $r$ of MCAR data $0$ (a), $10$ (b) and $30$ (c).}
\caption{Power simulation results $(\alpha=0.05)$ for $T^{\mu}$ $(\textendash\textendash\textendash)$, $T_{ML}$ ({\color{blue}$--$}), and $T_{tw}$ ({\color{ao(english)}$-\cdot-$}) for Symmetric and Asymmetric distributions under $\rho=0.5$ for testing $H_0^\mu:\{\mu_1=\mu_2\}$ with $n=10$ (left) and $n=20$ (right) and different percentages $r$ of MCAR data $0$ (a), $10$ (b) and $30$ (c).}
\caption{Low temperature gap dependence on kinetic energy for various values of $U$. Here we have set $\Delta t=4.5t$, $n=1.875$, $k_BT=0.01t$. The solid lines correspond to $V_{\rm SO}=0.5$. The dashed lines \red{shows the result for} $V_{\rm SO}=0$. A slightly larger value for the absolute value of the slope indicates that $V_{\rm SO}$ increases the asymmetry around the Fermi level.}
\caption{Critical temperature as a function of electron density for various values of the spin-orbit coupling with $\Delta t=0$. $U=-t$ (top left), $U=-2t$ (top right), $U=-3t$ (bottom left), and $U=-4t$ (bottom right). By particle-hole symmetry, the plot above half-filling is a reflection of this plot\red{, i.e. $T_c(2-n) = T_c(n)$ for $0 < n < 1$}.}
\caption{Critical temperature as a function of electron density for various values of the spin-orbit coupling with $\Delta t=4.5t$. \redd{ $U=90t$ (left), $U=115t$ (right).}}
\caption{Critical temperature as a function of electron density for various values of the spin-orbit coupling with $U=75t$. \redd{$\Delta t=3.5t$ (left), $\Delta t=4t$ (right).}}
\caption{Four types of cases in our results: 1). high relevance and correct answer; 2). low relevance and wrong answer; 3). high relevance but wrong answer; 4). low relevance but correct answer. ``(*,*)'' behind the explanations (attributes/caption) denotes the explanation-question relevance score and explanation accuracy, respectively. \colorbox[rgb]{ .682, .667, .667}{Gray} denotes groundtruth answers.}
\caption{Performance comparison with the state-of-the-art. We show the performance on both test-dev and test-standard splits of VQA-real open-ended task. The performances are achieved by training the VQA module on both the train and val splits. \colorbox[rgb]{ .682, .667, .667}{MCB-ensemble} incorporates external training data, which is an \textit{extra} advantage over other methods. \colorbox[rgb]{ .682, .667, .667}{Human} is the human performance for reference.}
\caption{Continuum subtracted ALMA spectrum for R~Dor around 336.8\,GHz in{\it black} extracted for a circular beam with aperture of 300\,mas. Two lines of previously identified molecules are observed: SO$_2$ ($\nu_2$=1, $20_{1,19}-19_{2,18}$), and TiO$_2$ ($23_{8,16} - 23_{7,17}$) indicated by the vertical dotted lines in {\it black}. The feature at 336.81603~GHz, tentatively identified as the FeO ($v$\,=\,0,$\Omega$\,=\,4,$J$\,=\, 11$-$10) transition, is blended with the line of TiO$_2$ at 336.82407~GHz. The flux of the TiO$_2$ line 8~MHz higher in frequency than FeO was estimated by referring to three lines of TiO$_2$ with similar excitation energies and channel maps that were observed in the same program (see inset in the right hand side of the panel). From these a \textit{\red{red} synthetic} profile for the TiO$_2$ line was derived. Plotted in {\it \color{blue} blue} is the observed profile minus the \textit{\red{synthetic}} TiO$_2$ profile attributed to FeO.}
\caption{Channel map of the FeO ($v$\,=\,0,$\Omega$\,=\,4,$J$\,=\,11--10) emission in R~Dor. The circle denotes the place of maximum dust emissivity (taking the contours at 1\%, 10\% and 90\% of the total flux). The local standard of rest velocity, $v_{\rm{LSR}}$, of R~Dor is $\sim$7\,km/s\citep{Decin2018}. The TiO$_2$ ($23_{8,16}-23_{7,17})$ transition at 266.8241\,GHz slightly blends the FeO line in the blue wing.}
\caption{ (a) The flux ($E >100$ MeV) versus photon spectral index of FRIs, FRIIs, SSRQs and NLSy1s. (b) Photon spectral index vs. \gray luminosity of the considered sources as compared with similar properties of blazars (light blue ellipse for FSRQs and light red for BL Lacs). In both plots FRIs, FRIIs, SSRQs and NLSY1s are shown with red, blue, cyan and green colors, respectively.}
\caption{Visual example for the computation of ternary convolutions without floating point operations. Ternary values are encoded by sign and value, i.e. +1 $\rightarrow$ (\textcolor{black}{$\blacksquare$},$\Box$), -1 $\rightarrow$ ($\Box$,$\Box$) and 0 $\rightarrow$ ($\Box$,\textcolor{yellow}{$\blacksquare$}). The approximation for a ternary filter bank provides scaling parameters $\alpha$ see below Eq.~\ref{eqTernaryWeight}. Ternary convolutions can be computed by masked XOR and XNOR operators followed by a bit-count according to Eq.~\ref{eqTernaryConv}. The output is batch normalised and passed on to the nonlinearity visualised in Fig.~\ref{figTHT}. }
\caption{Top row: A visual comparison of case \# 12 of the NIH demonstrates small but significant advantages of the ternary quantisation (middle) over the better performing adhoc binary activation and quantisation, which over-segments a neighbouring structure (left). Our approach better matches the manual segmentation (right). Bottom row: 3D visualisation of our segmentation shows a very smooth surface (left). Ranked (sorted) Dice score compared across methods demonstrate that the full-precision model is not significantly better than our heavily quantised TernaryNet. Both Binary XNORnet variants perform inferior.}
\caption{Indirect economic gains and losses of a 100- (\textcolor{blue}{blue}), 250- (\textcolor{red}{red}) and 1,500- (\textcolor{black}{black}) year flood event. Time labels on the x-axis indicate the end of each year, and the gray vertical bar marks the first year after the flood. The panels show the effects as percentage point changes relative to the baseline scenario, in which no disaster happens. (a) Cumulative changes in GDP-growth rates. (b) Cumulative changes in the unemployment rate. (c) Changes in the government debt-to-GDP ratio. Shaded areas cover one standard deviation above and below the mean values, as obtained from 50 independent Monte-Carlo simulations.}
\caption{Cumulative growth effects on sectoral GDP after a 250-year flood event for selected economic sectors in percentage points (pp) relative to the baseline scenario. Shaded areas cover one standard deviation above and below the mean values. Sectors shown: construction sector (black), manufacturing sector (\textcolor{red}{red}) and real estate sector (\textcolor{blue}{blue}). The gray vertical bar indicates the year of the flood.}
\caption{Cumulative changes in GDP growth relative to the baseline scenario as a function of the direct damage as a percentage of GDP. Results are shown for three different years after the disaster: \textcolor{blue}{2014}, \textcolor{red}{2015} and 2016. Shaded areas cover one standard deviation above and below the mean values. Immediately after the event (2014), all disaster sizes are associated with negative growth relative to the baseline scenario. In contrast, for the years 2015 and 2016 there exist inflection points and maxima for GDP growth, indicating the existence of direct damage sizes that, respectively, are ``optimal'' in terms of economic growth and determine a threshold where natural disasters become systemic events.}
\caption{Examples of classification results. We present the top-10 predicted categories and the corresponding probability scores. The ground-truth labels are highlighted in \textcolor{blue}{blue}. }
\caption{The iRobot Create\registered~ mobile robot used for all experiments. Some of the most important components are shown in red. On the top right of the image, the X-Y-Z coordinate system used is shown.}
\caption{NVE SD--MD simulations of a block of 500 fcc cobalt atoms with periodic boundary conditions, with initial conditions corresponding to a temperature of 300~K. The graphs plot the energy per atom and the norm of the total magnetization $|{\rm \bm{M}}|$ for timesteps sizes of $10^{-3}$~ps (black), $5\cdot 10^{-4}$~ps (red), and $10^{-4}$~ps (blue). The two insets plot the mean relative absolute deviations of the energy $\Delta {E}$ and magnetization $\Delta |{\rm M}|$ according to Eq.~\ref{relative_deviation} as a function of the timestep size $\Delta t$. The black dashed lines in the insets represent $ \mathcal{O} \left( \Delta t^2 \right)$ scaling behavior.} \label{fig:timestep_accuracy} \end{figure} Because the system is closed, it cannot adiabatically exchange energy with a surrounding reservoir and both quantities are very well conserved. Typical spin dynamics simulations use a timestep of $\Delta t = 10^{-4}$~ps. The current algorithm remains reasonably accurate up to a timestep size of $\Delta t = 10^{-2}$~ps. This demonstrates the strong stability of the integration algorithm. The influence of the timestep on the fluctuations in these quantities can be examined more precisely by calculating the mean relative absolute deviations averaged over the trajectory. For example, in the case of total energy per atom \begin{equation} \Delta {E} (\Delta t) = \frac{1}{N_{step}} \sum_{k=1}^{N_{step}} \left\lvert \frac{{E}_{k}(\Delta t) - \langle {E} \rangle (\Delta t) }{\langle {E} \rangle (\Delta t)} \right\rvert, \label{relative_deviation} \end{equation} where $E_{k}$ is the total energy per atom at timestep $k$ and $\langle{E} \rangle$ is the time average of the total energy per atom. %as a function of the number of timesteps, such as~: \begin{equation} %\langle {E} \rangle_{N} (\Delta t) = \frac{1}{N_{step}} %\sum_{k=1}^{N_{step}} {E}_{k}(\Delta %t) \label{timestep_average} \end{equation} The same formula can be applied to the norm of the total magnetization. The dependence of both of these quantities on the timestep size is shown in the two insets of Fig.~\ref{fig:timestep_accuracy}. The deviations increase as $\mathcal{O} \left(\Delta t^2 \right)$ . This behavior is consistent with the $\mathcal{O} \left( \Delta t^3 \right)$ accuracy of the ST decomposition, which describes propagation over a single step. For $N$ consecutive steps, the accuracy of the decomposition algorithm is given by $\mathcal{O} \left(N \Delta t^3 \right)$, with $N=t_{total}/\Delta t$. Thus, the accuracy for a fixed time interval is $\mathcal{O} \left( \Delta t^2 \right)$, as seen in Fig.~\ref{fig:timestep_accuracy}. For timestep size greater than $\Delta t = 10^{-2}$~ps, the scaling law starts to break down, indicating the the limit of numerical stability is being approached. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Spin--lattice dynamics in a canonical ensemble (NVT)}\label{sec:temperature} Now we connect the spin--lattice system to an external energy reservoir and stabilize the exchanges of energy with a given thermostat. In MD, many approaches have been already explored \cite{andersen_molecular_1980,nose_unified_1984,hoover_canonical_1985} and we focus on a Langevin strategy \cite{feller_constant_1995}. This is a widely used thermostat for atomic systems as well as coarse-grained models, e.g.~dissipative particle dynamics \cite{hoogerbrugge_simulating_1992} (see \cite{groot_dissipative_1997} for a review). In the Langevin approach, both random forces and damping terms are used in accord with the fluctuation-dissipation theorem. However, the presence of damping terms will lead to a contraction of the phase space volume, producing non-Hamiltonian dynamics. Yet, even in these cases, it has been shown that if some phase-space related measures are not preserved, pathological situations can be observed, and lead to incorrect distributions \cite{tuckerman_non-hamiltonian_2001}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Connecting spins to a random bath}\label{subsec1:temperature} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Néel first \cite{neel_influence_1949} and Brown later \cite{brown_thermal_1963} demonstrated that a sufficiently fine ferromagnetic particle consists of a single magnetic structure in which thermal agitation causes continual changes in the orientation of the moment. Such thermal fluctuations in the magnetization are well described by following the Langevin approach \cite{antropov_ab_1995,antropov_spin_1996,garcia-palacios_langevin-dynamics_1998}. In this approach, the spin system is connected to a single thermal bath modeled by an infinite number of degrees of freedom. The properties of such bath are given by $\bm{\eta}$, a random vector, whose components follow a Gaussian probability law given by the first and second moment \begin{eqnarray} \langle {\bm\eta}(t) \rangle &=& {\bm 0} \nonumber \\ \langle \eta_{\alpha}(t) \eta_{\beta}(t') \rangle &=& 2 D_S \delta_{\alpha\beta}\delta (t-t') \label{white_noise} \end{eqnarray} where $\alpha$ and $\beta$ are the vector components. In order to preserve the norm of each individual spin, this random fluctuation is added in a multiplicative manner to Eq.~\ref{LLG} via a random torque \cite{mayergoyz_nonlinear_2009}, that leads to the following stochastic Landau-Lifshitz-Gilbert equation~: \begin{equation} \frac{d \bm{s}_{i}}{dt} = \frac{1}{\left(1+\lambda^2 \right)} \left( \left(\bm{\omega}_{i} +\bm{\eta} \right) \times \bm{s}_{i} + \lambda\, \bm{s}_{i}\times\left( \bm{\omega}_{i} \times\bm{s}_{i} \right) \right). \label{sLLG} \end{equation} From Eq.~\ref{sLLG}, one can derive a Fokker-Planck (FP) equation \cite{garcia-palacios_langevin-dynamics_1998}. The resolution at equilibrium of the FP equation allows a fluctuation--dissipation relation to be derived that assigns a proportion of the amplitude of the noise $D_S$ to the given external thermostat $T$~\cite{mayergoyz_nonlinear_2009}, such that \begin{equation} D_S = \frac{2 \pi \lambda k_B {T} }{\hbar}, \label{spin_fluctuation_dissipation} \end{equation} with $k_B$ = the Boltzmann constant. A subtle detail about stochastic differential equations is a proper choice of a stochastic prescription that allows an evaluation of the ill-defined random vector~: \begin{equation} \bm{\mathcal{I}}(\Delta t) ~=~ \int_{t}^{t+\Delta t} \bm{\eta}(t') \times \bm{s}_{i}(t') dt', \end{equation} by defining a point within the interval $[t, t+\Delta t]$ at which this integral is approximated consistently. Numerous papers have discussed various choices of the prescription for stochastic magnetization dynamics performed with the stochastic LLG equation \cite{garcia-palacios_langevin-dynamics_1998,daquino_midpoint_2006,aron_magnetization_2014}. We use the mid--point Stratonovich approach that preserves the norm of individual spins and has time micro-reversibility. %%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Measuring lattice and spin temperatures}\label{subsec3:temperature} %%%%%%%%%%%%%%%%%%%%%%%%%%% In non-equilibrium spin-lattice systems, it is convenient to use a single thermostat to control the flow of entropy production between the system and a thermal reservoir \cite{rapaport_art_2004}. Such a thermostat has to be related to a microcanonical quantity defined in a statistical ensemble that measures the temperature. Because the temperature of an equilibrium system is calculated from the mean kinetic energy of its particles, the transient kinetic energy for spins is not an obvious quantity. An immediate consequence is how to measure the temperature $T_{S}$ that controls the thermostatting of the magnetic degrees of freedom \cite{antropov_spin_1996}. At equilibrium, the instantaneous lattice temperature is usually defined as the kinetic temperature of the atoms: \begin{equation} T_{L} = \frac{2}{3\, N\, k_B} \sum_{i=1}^{N} \frac{|\bm{p}_{ i}|^2}{2 m_{i}}. \label{lattice_temperature} \end{equation} This expression of the kinetic temperature is known to rely on some approximations, such as the equivalence of ensembles and the ergodicity assumption \cite{rugh_dynamical_1997}. Other dynamical approaches for measuring the temperature in Hamiltonian systems within a microcanonical ensemble have been derived, and proved to be accurate \cite{rugh_geometric_1998,jepps_microscopic_2000}. However, we will assume that the kinetic expression given by eq.~\ref{lattice_temperature} is sufficient for the purpose of this work. %(NOTE: the next sentence does not make sense to me. Also these next 2 paragraphs are somewhat unclear. I'm not sure why you need to debate how to compute the atom temp or spin temp. You could just use eq (22) and be done with it.) We first ignore the number of degrees of freedom, frozen by the instantaneous determination of the local streaming velocity. Second, it assumes the thermodynamical temperature, given by the variation of the internalenergy respectively to the entropy in a non equilibrium system is identical to the kinetic temperature. That is true for a lattice system only by the additional postulate that the local thermodynamic equilibrium is exact. Thus the thermodynamic temperature can be assumed to be a monotonic function of the kinetic temperature only, even outside the linear regime, when a unique non-equilibrium steady state can be reached by the system. %Because temperature and entropy are thermodynamic conjugate variables, Rugh \cite{rugh_dynamical_1997,rugh_geometric_1998} has proposed a method of determining the temperature in a Hamiltonian system by considering the entropy to be a canonical invariant weighted area of the energy surface, under the assumption that the dynamical system is ergodic on the energy surface. Thus arbitrary phase space vector fields can be used to generate phase functions whose ensemble averages give the thermodynamic lattice temperature \cite{jepps_microscopic_2000}. %As already demonstrated for Hamiltonian systems of Newtonian particles \cite{butler_configurational_1998}, this enables derivation of an expression for the temperature of the spin system. Transcripting Rugh's the geometrical approach to spin systems \cite{rugh_dynamical_1997}, and when the thermodynamic limit is considered, Nurdin \emph{et al.} \cite{nurdin_dynamical_2000} give a spin temperature as \begin{equation} T_{S} = \frac{\hbar}{2 k_B} \frac{\sum_{i=1}^{N} |\bm{s}_{i} \times \bm{\omega_{i}}|^2}{\sum_{i=1}^{N} \bm{s}_{i} \cdot {\bm\omega}_{i}} \label{spin_temperature} \end{equation} Another approach was later derived by Ma \emph{et al.} \cite{ma_temperature_2010}. Relying on the fluctuation-dissipation theorem, it gave an analog definition of the temperature of a spin ensemble. In refs.~\cite{beaujouan_anisotropic_2012,perera_reinventing_2016}, the definition given by eq.~\ref{spin_temperature} has proven to be efficient for the thermalization of the spin subsystem and its relaxation toward thermal equilibrium during spin--lattice simulations. Besides, this definition of $T_{S}$ has the same domain of validity as the kinetic temperature expressions for $T_{L}$ defined above. Thus, we chose to use the definition of the spin temperature defined by Nurdin \emph{et al.}. %%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Thermalizing the spin--lattice system}\label{subsec2:temperature} %%%%%%%%%%%%%%%%%%%%%%%%%% For the simulation of relaxation processes, the connecton of the lattice system to a thermal bath is also performed using the Langevin approach. Eq.~\ref{s_advance} is replaced by Eq.~\ref{sLLG}, and a new random force and a damping term are added to Eq.~\ref{p_advance}. This yields the following equations that model stochastic magnetic molecular dynamics: \begin{eqnarray} \frac{d\bm{r}_{i}}{dt} &=& \frac{\bm{p}_{i}}{m_{ i}} \label{r_advance_T} \\ \frac{d\bm{p}_{i}}{dt} &=& \sum_{j,i\neq j}^{N}\left[-\frac{dV\left(r_{ ij}\right)}{dr_{ij}}+ \frac{dJ\left( r_{ ij}\right)}{dr_{ij}} \bm{s}_{i} \cdot \bm{s}_{j} \right] \bm{e}_{ij} \nonumber \\ ~&~& -\frac{\gamma_L}{m_{ i}} \bm{p}_{i} +\bm{f}(t) \label{p_advance_T}\\ \frac{d\bm{s}_{i}}{dt} &=& \frac{1}{\left(1+\lambda^2 \right)}\left(\left(\bm{\omega}_{i} +\bm{\eta}(t)\right) \times \bm{s}_{i} + \lambda\, \bm{s}_{ i}\times\left( \bm{\omega}_{i} \times\bm{s}_{i} \right) \right) \label{s_advance_T} \end{eqnarray} In Eq.~\ref{p_advance_T}, $\gamma_L$ is a damping parameter, and $\bm{f}$ its corresponding fluctuating force drawn from a Gaussian distribution with \begin{eqnarray} \langle {\bm f}(t) \rangle &=& {\bm 0} \\ \langle f_{\alpha}(t) f_{\beta}(t') \rangle &=& D_L \, \delta_{\alpha\beta}\, \delta(t-t') \end{eqnarray} where $\alpha$ and $\beta$ are coordinates, and $D_L$ is the amplitude of the random variables. $D_L$ can be parametrized according to the fluctuation--dissipation relation, and then depends on the temperature of the thermal bath coupled to the lattice, and on the damping coefficient $\gamma_L$ according to the Einstein relationship \cite{tuckerman_statistical_2010}. The probability distribution of the noise vector $\bm{\eta}$ follows Eqs.~\ref{white_noise} and \ref{spin_fluctuation_dissipation} respectively. Extensions to more than two out-of-equilibrium dynamics (here referring to spin and lattice dynamics) are possible and several authors linked the damping coefficients to spin--electron and lattice--electron relaxation processes \cite{ma_spin-lattice-electron_2012}. This allowed using the model presented above to simulate spin--lattice--electron relaxation processes that occur in ultrafast magnetic switching experiments. However there are concerns about the choice of the noise correlation functions. In this study, we decided to remain within the framework of the Markov hypothesis, and focus on uncorrelated white-noise only. When the characteristic timescales of the dynamics reach values as small as those of the simulated relaxation processes, the Markov hypothesis may break down and a colored-noise, such as an Ornstein–Uhlenbeck process, becomes a better approximation of the exchange of causal information between the system and its reservoir. Recent studies have focused on evaluating the influence of such memory effects on the magnetization dynamics \cite{atxitia_ultrafast_2009,bose_correlation_2010,tranchida_closing_2016,tranchida_colored-noise_2016}, and suggest non-trivial magnetization dynamics beyond the second order cumulant of the spin variables. These points will be addressed in a separate study. In order to evaluate the efficiency of the two-thermostat model presented here, two different simulations were performed. A cell of 500 cobalt atoms on an fcc lattice, coupled by three interactions (the magnetic exchange interaction, a spin-orbit coupling, and a mechanical EAM potential \cite{thibaudeau_thermostatting_2012,pun_embedded-atom_2012}) is considered. For the first simulation, a Langevin thermostat was applied only to the spins, according to Eq.~\ref{s_advance_T}. The simulation started from a fixed lattice equilibrium configuration, corresponding to $T_{L}=0$K and an initial spin configuration sampled from a $T_{S}=300$K magnetic equilibrium state. The second simulation was exactly the opposite: the Langevin thermostat was applied only to the motion of the lattice atoms, according to Eq.~\ref{p_advance_T}. The simulation started in a configuration with all the spins aligned along their effective fields, which corresponds to $T_{S}=0$K, and with atom velocities sampled from a $T_{L}=300$K equilibrium lattice state. Fig.~\ref{fig:spinlatt_relaxation} plots the time evolution of $T_{ L}$ and $T_{S}$ for the two simulations. \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{NVT_spin_latt_relaxation.pdf} \caption{Measure of the spin T$_{S}$ and lattice T$_{L}$ temperatures during two NVT simulations of 500 cobalt atoms. For both graphs, the black dashed lines plot the target temperature T$_{T}$, and the green and red curves plot the lattice and spin temperatures, T$_{L}$ and T$_{S}$, respectively. In the upper graph, the Langevin thermostat was applied to the spins only, according to Eq.~\ref{s_advance_T}, with a target T$_{T}$=300K, and the transverse magnetic damping set to $\lambda = 0.01$. In the lower graph, the thermostat was applied only to the atomic motion, according to Eq.~\ref{p_advance_T} with T$_{T}$=300K and a lattice damping parameter of $\gamma_L=10$ s$^{-1}$.} \label{fig:spinlatt_relaxation} \end{figure} In each case, the thermostatted degrees of freedom stayed at the target temperature for the duration of the simulation. And the non-thermostated degrees of freedom relaxed from their initial temperature to the thermostatted temperature due to the spin--lattice coupling of the dynamics. In both cases the relaxation time was a few 100 ps, which is consistent with the spin--lattice coupling time shown in Fig.~\ref{fig:3T}. This illustrates how proper choice of damping coefficients can produce physically correct responses for magnetic exchange, spin-orbit, and mechanical coupling interactions. To better understand how the magnetic and mechanical energy couple to each other, a larger NVE simulation was performed with 1372 cobalt atoms initially on an fcc lattice. The initial spin configuration was random so that the spin system is in a paramagnetic state at a very high temperature. The atomic velocities (lattice temperature) were initialized to 200K. The potential energy of the atoms was initially at a minimum due to the perfect lattice configuration. As the simulation evolved energy re-partitioned between these 3 components (kinetic energy, potential energies of atoms and spins) but the total energy remained constant with no perceptible fluctuations (at this scale) as seen in the upper graph of Fig.~\ref{fig:energies}. The lower graph of Fig.~\ref{fig:energies} plots the evovling spin and lattice temperatures, which equilibrated to the same value, indicating the efficacy of the spin--lattice coupling. \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{NVE_energy_relaxation.pdf} \caption{NVE simulation of a periodic block of 1372 cobalt atoms. Spin directions are initially random. Initial particle velocities were drawn from a Maxwell distribution at 200K. The upper graph plots 3 contributions to the total energy which remained constant: kinetic energy of the atoms and potential energies of the atoms and spins (magnetic energy). The lower graph plots the atomic (lattice) and spin temperatures, corresponding to Eqs.~\ref{lattice_temperature} and \ref{spin_temperature} respectively, which equilibrated to the same value.} \label{fig:energies} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Parallel implementation of the spin--lattice algorithm}\label{sec:parallel} \subsection{Synchronous sub-lattice algoritm}\label{subsec1:parallel} To exploit the power of parallel processing, we developed a spatially-decomposed version of the serial SD--MD algorithm described in Section~\ref{sec:algorithms}, analogous to that used for molecular dynamics simulations in LAMMPS. Like the serial algorithm, it is essential that the parallel algorithm accurately conserve energy and the norm of the total magnetization. As explained in Section~\ref{sec:algorithms}, this requires that the spin propagation operators be applied in a manner that preserves symplecticity. This in turn requires that the magnetic force ${\boldsymbol \omega}_i$ acting on each spin ${\boldsymbol s}_i$ be calculated when each spin is updated. Due to short-range spin-spin interactions, the computation of this force depends on the spin orientations of the neighbors of atom $i$. In other words, before updating each spin, the current value of its neighboring spins must be known. The parallel issue arises when two neighbor spins ${\boldsymbol s}_i$ and ${\boldsymbol s}_j$ (and their associated atoms) reside on different processors. How do we ensure that when the second spin is updated, it uses the previously updated value of the first spin? This clearly requires some inter-processor communication of spin information during the update operation, but the cost of the communication needs to be minimized to ensure an efficient algorithm. Likewise we must ensure two neighbor spins on different processors are not updated simultaneously (one of them with outdated information). Otherwise the accuracy of the integration scheme will be degraded.. Inspired by the lattice-dependent ST decomposition proposed by Krech \emph{et al.}~\cite{krech_fast_1998}, Ma \emph{et al.} \cite{ma_parallel_2009} developed a multithreading algorithm for spin dynamics that partitions the single-spin evolution operators into groups whose member spins do not interact. This allows all spins in different groups to be updated concurrently by separate execution threads. The multi-threading implementation delivers good speedup on a single multicore CPU. As currently implemented, distributed-memory parallel execution is not supported, although the method could be extended to use spatial decomposition parallel algorithms. The biggest limitation of the method is the lack of generality. Partitioning into non-interacting groups is achieved using a lattice-coloring or checkerboarding scheme that can only be applied to perfectly regular lattices in which each spin interacts with a fixed stencil of interacting neighbor spins. Moreover, as the number of neighbors in the interaction stencil increases, the number of groups required to achieve a correct ST decomposition also increases, adding to the complexity of the algorithm. We have implemented a different method called sectoring which is based on the synchronous sub-lattice algorithm \cite{shim_semirigorous_2005, plimpton_crossing_2009} used to achieve spatial parallelism in kinetic Monte Carlo simulations. In contrast to checkerboarding, sectoring is very general, requiring only that spin-spin interactions vanish beyond a finite cutoff distance. As long as this requirement is met, sectoring can be used for dynamic simulations of arbitrary systems of atomic spins, including perfectly regular lattices, thermally vibrating lattices, spatially disordered systems, and even systems undergoing diffusive dynamics in which the set of neighbors of each spin can change over time. The sectoring method can be thought of as an extension to the spatial-decomposition MD algorithm used in LAMMPS and other MD codes, where the system is partitioned into subdomains, one per processor. Each processor owns and time-integrates the atoms in its subdomain. It also stores information about nearby ghost atoms, up to a cutoff distance away, which are owned by neighboring processors. This information is acquired by inter-processor communication, when needed. The sectoring idea is to further divide each subdomain into smaller regions by bisecting the subdomain once in each physical dimension (4 sectors in 2d, 8 in 3d). If the spatial extent of all sectors in any dimension is larger than the interaction cutoff distance, then spins in the same sector on two different processors do not interact. The processors can thus concurrently update all the spins in one sector without the need to communicate spin information to/from other processors, while still adhering to the ST decomposition of Eq. \ref{spin_ST_decomposition}. Formally, we can rewrite the ST decomposition as \begin{equation} e^{\hat{L}_{s}\frac{\Delta t}{2}} = \prod_{{ k}=1}^{K} \left \{ \prod_{{ j}=1}^{N_k} e^{\hat{L}_{s_{j}}\frac{\Delta t}{4}} \right \} \prod_{{ k}={K}}^1 \left \{ \prod_{{ j=N_k}}^{1} e^{\hat{L}_{s_{j}} \frac{\Delta t}{4}} \right \} +\mathcal{O}\left( \Delta t^3 \right), \label{spin_ST_sectoring} \end{equation} where ${K}$ is the number of sectors, ${N_k}$ is the number of spins in sector ${k}$, and ${s_{j}}$ is the ${j}$th spin in the sector. \begin{figure}[ht] \centering \includestandalone[width=0.5\columnwidth]{figuresector} \caption{Schematic of the SD--MD sectoring algorithm for a two--dimensional spin--lattice model. The system is divided into four subdomains (bold squares), each owned by a different processor. Each subdomain is further divided into four sectors (numbered squares). Spins in sector 1 (shaded) for two different processors do not interact and can thus be updated concurrently without interprocessor communication.} \label{fig:sectoring} \end{figure} Fig.~\ref{fig:sectoring} illustrates this idea for a two-dimensional system running on 4 processors, with each processor subdomain further divided into 4 sectors. Spins in sector 1 of a particular processor do not interact with spins in sector 1 of any other processor. Thus all the processors can update their sector 1 spins at the same time. After the sector 1 updates, new spin values must be communicated between processors before sector 2 spins can be updated. Fig.~\ref{fig:communication} is a schematic of the necessary communication for the same 2d system. Each processor sends the current values of spins associated with the ghost atoms that border sector 1. The thickness of the border region is equal to the spin-spin interaction cutoff distance. Note that Fig.~\ref{fig:communication} shows the minimal communication requirements. For simplicity and convenience, our current implementation makes use of standard communication functions in LAMMPS. All the ghost spins adjacent to the subdomain are updated, not only those adjacent to sector 1. This simplifies the implementation at the expense of a certain amount of unnecessary updating of spins that have not changed value. However, in either case, the cost of computation scales as $O((N/P)^{2/3})$, with $N$ the number of atoms and $P$ the number of MPI processes, differing only in the prefactor\cite{plimpton1995fast}. In applications where the communication cost is large relative to computation, extra overall performance could be achieved by limiting the interprocessor communication to only those sites adjacent to sector 1. This pattern of communicating ghost spins then updating a sector is repeated four times, once for each sector (8 times in 3 dimensions). The entire process is then repeated in reverse sector order, at which point all of the spins have been updated by a half timestep, according to Eq. \ref{spin_ST_sectoring}. Overall, spins in each sector are sequentially updated four times per timestep. To accurately integrate the spin dynamics, it is important that a fixed ordering of spins be used for all four updates. We achieve this by assigning atoms to sectors once at the start of the timestep based on their current positions. %(NOTE: there are actually 2 cylces of forward and backward thru %the sectors that are needed, for just 1/2 a timestep? So %there are 4 cycles thru the sectors per timestep?) %(NOTE: I think Fig 6 should be changed so that the dashed lines %are around the sector 2 owned by the the upper-left processor %in Fig 5. So that there is no confusion about periodic boundaries %in the schematic, i.e. everything is internal.) %(NOTE: is there some requirement that the forward and reverse ordering %of atoms (within each sector) on the same timestep need to be %identical? And that is why you use the per-sector linked lists, to %store these orders? If so, that should be mentioned.) \begin{figure}[ht] \centering \includestandalone[width=0.5\columnwidth]{figurecomm} \caption{Schematic of the communication required before a processor can update spins in sector 1. This is a zoomed-in view of the bottom left portion of Fig.~\ref{fig:sectoring}. The current values of all spins within the dashed square must be known before the processor updates spins in sector 1 (shaded). Most of these values are already known, because the spins are owned by the processor. But ghost spins in the blue regions are owned by neighboring processors, Those other processors must the send the current values to this processor. Similarly, this processor must send the current values of its spins in the red regions to neighboring processors. Once these communication operations have been completed, all the processors can concurrently advance their spins in sector 1. } \label{fig:communication} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Accuracy of the parallel algorithm}\label{subsec2:accuracy} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Note that by construction, the parallel algorithm faithfully reproduces the ST decomposition rule that each spin is updated using current information for all its neighbor spins. It differs from the serial algorithm only in the order in which the global set of spins are updated. This order will also differ depending on the number of processors used to run a simulation. Two simulations with different ordering will not evolve identically in a numerical sense, but they should be produce identical in a statistical sense, since both are ``correct''. We tested this using the same fcc cobalt model used for testing the accuracy of the serial algorithm (Fig.~\ref{fig:timestep_accuracy}). To allow for multiple processors, the size of the periodic simulation cell was increased from 17.7~\AA~(500 atoms) to 28.32~\AA~(2048 atoms). For a $2 \times 2 \times 2$ grid of 8 processors, the width of each sector is 7.08 \AA, which is larger than the cutoff distance of 6.5~\AA~used by the EAM potential, and the one of 4.0~\AA~of the magnetic interactions. %(NOTE: is the EAM cutoff the same as the spin/spin cutoff?) Starting with the same configuration initialized at 300~K, we ran serial and parallel NVE SD--MD simulations for 1~ps using a $10^{-4}$ps timestep. Fig.~\ref{fig:accuracy_serial_vs_mpi} plots the conservation of total energy and magnetization for the serial algorithm as well as the parallel algorithm running on 4 and 8 MPI processes. \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{mpi_accuracy.pdf} \caption{NVE SD--MD simulation of a block of 2048 fcc cobalt atoms using the sectoring parallel algorithm. The upper graph plots the energy E(t) minus the initial energy E$_0$ as a function of simulation time. The lower graph plots the norm of the total magnetization $|{\rm \bm{M}}|$ minus the initial value $|{\rm \bm{M}}|_0$. The black curves are for the serial algorithm, and the red and green curves (or dots for the upper graph) are for the parallel algorithm.} \label{fig:accuracy_serial_vs_mpi} \end{figure} The upper graph of Fig.~\ref{fig:accuracy_serial_vs_mpi} shows the serial and parallel algorithms generate trajectories with tiny energy fluctuations that are indistinguishable from each other (the red and black curves overlay), especially relative to the size of the step-to-step variations in energy due to the numerical integration. The lower graph of Fig.~\ref{fig:accuracy_serial_vs_mpi} shows there there are tiny differences between the simulations in the total magnetization value at a particular timestep. This is expected, because the spins are updated in a different order in each simulation, leading to trajectories that diverge in a numerical sense. However the overall step-to-step variation in the magnetization is the same for all three simulations, indicating the parallel algorithm is generating a spin trajectory that is statistically equivalent to that of the serial simulation. %%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Scaling results}\label{subsec3:parallel} %%%%%%%%%%%%%%%%%%%%%%%%%% We now evaluate the scaling efficiency of the parallel spin--lattice algorithm as implemented in LAMMPS, for both strong and weak scaling. All simulations were performed on a cluster at Sandia consisting of dual-socket Intel Xeon E5-2695 (Broadwell) CPUs with 36 cores per node, and an Intel Omnipath interconnect. To ensure each processor owns the same number of spins, only 32 cores per node were used for each of the calculations, though this is not required for general runs. Both strong and weak scaling tests were performed. In both cases, the parallel efficiency was evaluated by computing a normalized simulation rate (SR), defined as the number of atoms - steps per CPU second and per node: \begin{equation} {\rm SR} = \frac{{\rm Steps} \times {\rm Atoms} } {{\rm Nodes} \times {\rm Seconds} } \label{pareff} \end{equation} and plotted as a function of the number of nodes. Strong scaling was tested by increasing the number of nodes for a fixed number of atoms. Ideally, the computation time should be cut in half each time the number of nodes is doubled, so that SR should remain constant. %fixed size simulation, measuring the reduction in computation time. By %comparing the results to the computation time for the same problem %treated with the serial algorithm, and by checking how the results are %diverging from ideal scaling, we can evaluate the efficiency of the %parallel implementation as the amount of work per process gets smaller %and smaller. Each NVE SD--MD simulation combined a mechanical EAM potential with two magnetic interactions, the exchange energy and a N\'eel anisotropy (see \ref{app:exchange} and \ref{subapp1:soc} for more details). The magnetic potentials were parametrized according to Ref.~\cite{beaujouan_anisotropic_2012}; the EAM potential is described in Ref.~\cite{pun_embedded-atom_2012}. Two different problem sizes were run, the smaller with 256,000 (256K) atoms (a cubic box of $40^3$ fcc unit cells, 4 atoms/cell) and the larger with 2,048,000 (2M) atoms (2x larger in each dimension). The SR defined in Eq.~\ref{pareff} was averaged over a 50 timestep run. To assess the cost and efficiency of the SD--MD algorithm, the same simulations were run with only the EAM potential (no spin variables and without the sectoring algorithm). Fig.~\ref{fig:strongscaling} shows the results. \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{strong_scaling_nodes.pdf} \caption{Strong-scaling performance of the parallel SD--MD algorithm (red) compared to standard MD (blue) for two fixed size problems. The smaller system (circles) has 256,000 atoms, the larger (squares) has 2,048,000 atoms. Perfect scaling would be horizontal lines.} \label{fig:strongscaling} \end{figure} Fig.~\ref{fig:strongscaling} indicates the overall computational cost of including magnetic spin interactions, including the 8x increase in inter-processor communication to perform the sectoring algorithm, is about 4-5x that of a standard MD simulation with EAM potentials. %(NOTE: is it really 8x more comm, or not that much?) For perfect strong scaling, the SR should remain constant as the number of nodes increases. As the smaller system is run on more nodes, the relative cost of interprocessor communication grows as the atoms/node ratio shrinks, and the SR decreases slowly. The effect is less pronounced for the larger problem. Note that for the smaller problem, the maximum node count was 16 nodes ($32\times 16=512$ MPI tasks) to satisfy the requirement that the sector width be at least equal to the spin-spin interaction cutoff distance (4~\AA). Also note that the higher computational cost of the SD--MD model results in somewhat better strong scaling efficiencies relative to standard MD. Weak scaling was tested by increasing the size of the system as the number of nodes was increased, so that the number of atoms per node (and thus the size of each processor sub-domain) was held fixed. The same magnetic and EAM interaction models were used as for the strong scaling runs. Two different sized systems with 32,000 (32K) (corresponding to 1,000 atoms per processor), and 256,000 (256K) atoms/node (corresponding to 8,000 atoms per processor) were simulated. Fig.~\ref{fig:weakscaling} shows the results for a number of nodes ranging from 1 to 64 (corresponding to 2,048 processes). \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{weak_scaling_nodes.pdf} \caption{Performance of the parallel SD--MD algorithm compared to the standard MD algorithm for scaled size problems (weak scaling). In both cases, the interaction models are the same as that used in Fig.~\ref{fig:strongscaling}. The simulation rate (SR in Eq.~\ref{pareff}) is plotted as a function of the number of nodes. SD--MD and MD results are colored red and blue, respectively. The smaller scaled systems (circles) consisted of 32,000 atoms/node and the larger scaled systems (squares) consisted of 256,000 atoms/node. } \label{fig:weakscaling} \end{figure} For ideal weak scaling, the value of SR should remain constant as the number of nodes increases, because the amount of work per process remains constant. Fig.~\ref{fig:weakscaling} shows that SR indeed decreases very little, confirming the efficiency of our SD--MD algorithm. Also, we observe that our SD--MD algorithm is again approximately 4 to 5 times slower than the standard NVE MD calculation with LAMMPS, with the same EAM mechanical potential. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Conclusion}\label{sec:conclusion} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% A parallel implementation of coupled magnetic and molecular dynamics was presented. This implementation of SD--MD is available as a new package in LAMMPS. It allows both coupled and uncoupled simulations to be performed, i.e. magnetic spins on a fixed lattice of atoms or on mobile atoms. Both the serial and parallel versions implement statistically equivalent symplectic Suzuki-Trotter decompositions of the spin propagation operators, differing only in the order they choose to update spins. The accuracy of the serial version of the coupled algorithm was analyzed in both the microcanonical (NVE) and canonical (NVT) statistical ensembles. In the NVE case, care was taken to verify that both the total internal energy and the norm of the magnetization were conserved up to second order in the timestep size. For NVT simulations testing coupled spin--lattice relaxation, Langevin thermostats were applied to either the spin subsystem or the lattice subsystem. A parallel implementation of the integration algorithm based on the sectoring or synchronous sublattice method was described in detail. The statistical equivalence of the parallel algorithm to the serial algorithm was verified by monitoring the conservation of total energy and the norm of the magnetization as the number of parallel processors varied. The parallel performance of the LAMMPS implementation was assessed, both for fixed size problems (strong scaling) and scaled size problems (weak scaling). In both cases, good performance was observed for all cases with more than 500 atoms per processor. Our new SD--MD algorithm with coupled magnetic spin and atomic lattice dynamics was shown to be only 4 to 5 times slower than an analogous NVE MD LAMMPS run treating only the motion of the atomic lattice. This sectoring method has the advantage of being very general, working for both perfectly ordered particle configurations and disordered systems, so long as the width of each processor domain is at least twice the cutoff distance for the short-range spin-spin interactions. Because the new methods were implemented within the open-source LAMMPS MD code, they are now available to the scientific community, enabling a wide variety of coupled spin--lattice simulations to be easily performed and tested in detail, as well as new models and algorithms to be implemented. Applications of this method to the simulation of magnetoelastic effects in magnetic alloys will be presented in subsequent publications. \section*{Acknowledgement} Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia LLC, a wholly owned subsidiary of Honeywell International Inc. for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-NA0003525. J.T. acknowledges financial support through a joint CEA--NNSA fellowship. J.T. would also like to thank L. Berger-Vergiat and L. Bertagna for fruitful conversations, and S. Moore for interesting comments and his reviews of this work. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\section*{Notes} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\begin{itemize} %\item Try to work in SPPARKS reference with Shim reference %\item emphasize symplecticity %\item check for consistent use of terms SD--MD, sector, domain %\end{itemize} %% The Appendices part is started with the command \appendix; %% appendix sections are then done as normal sections %% \appendix %% \section{} %% \label{} %% References %% %% Following citation commands can be used in the body text: %% Usage of\cite is as follows: %%\cite{key} ==>> [#] %%\cite[chap. 2]{key} ==>> [#, chap. 2] %%\citet{key} ==>> Author [#] %% References with bibTeX database: %\nocite{*} \bibliographystyle{model1-num-names} \bibliography{MDSD_redaction} %% Authors are advised to submit their bibtex database files. They are %% requested to list a bibtex style file in the manuscript if they do %% not want to use model1-num-names.bst. %% References without bibTeX database: % \begin{thebibliography}{00} %% \bibitem must have the following form: %% \bibitem{key}... %% % \bibitem{} % \end{thebibliography} \appendix %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Definition of the spin Hamiltonian}\label{app:hamiltonian} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In the equations of motion (eq.\ref{p_advance_T}), the mechanical force coming from the spins can be computed from the partial derivation of a spin Hamiltonian. This Hamiltonian represents the total energy of the magnetic systems. Its general form can be separated according to the following expression~: \begin{equation} \mathcal{H} = \mathcal{H}_{z} + \mathcal{H}_{ex}+ \mathcal{H}_{an}+ \mathcal{H}_{N\acute{e}el}+ \mathcal{H}_{dm}+ \mathcal{H}_{me} +\mathcal{H}_{di}, \end{equation} with $\mathcal{H}_{z}$ and $\mathcal{H}_{ex}$ the Zeeman and the exchange interactions defined in Section~\ref{sec:dynamics}, the magnetic anisotropy (2 terms), the Dzyaloshinskii-Moriya, the magneto-electric, and the dipolar interaction. The subsections of this appendix give a definition of these interactions. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Parametrization of the exchange interaction}\label{app:exchange} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Because of the celebrated Bohr-van Leuween theorem \cite{van_leeuwen_problemes_1921}, the flow of permanent magnetic moments cannot be the origin of the magnetism found in actual materials and more surprisingly, magnetism is an inherently quantum mechanical effect. It is the interplay of electronic properties which appear unrelated to magnetism, the Pauli principle in combination with the Coulomb repulsion as well as the hopping of electrons that leads to an effective coupling between the magnetic moments in a solid \cite{duck_pauli_1998}. This mechanism does not have a classical analogue and is responsible for both the volume of matter and ferromagnetism. In SD-MD, an effective parametrization of this exchange interaction comes with two consequences : First, it is assuming that its intensity is rapidly decreasing with very few oscillations of its sign, so that a rigid cutoff radius $R_{c}$ can be safely introduced. For the computation of the exchange interaction with a given spin $i$, only spins $j$ such that $r_{ij} < R_{c}$ have to be taken into account. Second, the value of its intensity can be approximated by a simple continuous isotropic function ${J}\left(r_{ij} \right)$. For $3d$ atoms, this function is based on the Bethe--Slater curve \cite{kaneyoshi_introduction_1992, yosida_theory_1996}, and is parametrized via three coefficients, $\alpha$ in eV, $\delta$ in $\angstrom$, and $\gamma$ a non-dimensional constant. One has~: \begin{equation}{J}\left(r_{ij} \right)= 4 \alpha \left(\frac{r_{ij}}{\delta} \right)^2 \left(1 - \gamma \left(\frac{r_{ij}}{\delta} \right)^2 \right)e^{-\left(\frac{r_{ij}}{\delta} \right)^2 }\Theta (R_c - r_{ij}), \label{BSfunction} \end{equation} with $\Theta (R_c - r_{ij})$ the Heaviside step function. \begin{figure}[ht] \centering \includegraphics[width=0.95\columnwidth]{exchange_graph.pdf} \caption{Examples of exchange curves for bcc-Fe in blue, fcc-Ni in red, fcc-Co in green, and bismuth ferrite oxide (BFO) in purple. The dotted line labeled R$_{c}$ represents a cutoff radius taking into account the three first neighbor shells.} \label{fig:exchange} \end{figure} As examples, Fig.~\ref{fig:exchange} plots the interpolation by the function presented in Eq.~\ref{BSfunction} of different exchange data. On one side, the three first sets are related to pure ferromagnetic metals such as Iron, Nickel and Cobalt, and come from \emph{ab initio} calculations performed within the non-relativistic spin-polarized Green-function technique \cite{pajda_ab_2001}, in general agreement with scattering experiments. On the other side, the fourth set is for anti-ferromagnetic BiFeO$_3$ (BFO) and was measured by inelastic neutron scattering technique \cite{jeong_spin_2012}. The values of the coefficients $\alpha$, $\gamma$ and $\delta$ to use for each of these four materials are gathered in table~\ref{tab:exchange}. \begin{table}[ht] \begin{center} \begin{tabular}{|c|c|c|c|} \hline ~ & $\alpha$ (meV) & $\gamma$& $\delta$ ($\angstrom$) \\ \hline \hline Bcc-Fe & 25.498 & 0.281 & 1.999 \\ \hline Fcc-Ni & 9.73 & $1.08\cdot 10^{-4}$ & 1.233 \\ \hline Fcc-Co & 22.213 & $8.08\cdot 10^{-6}$ & 1.485 \\ \hline BFO & -15.75 & 0 & 1.965 \\ \hline \end{tabular} \caption{Coefficients fitting the \emph{ab initio} and experimental results (with the model presented in Eq.~\ref{BSfunction}) for bcc-Fe, fcc-Ni, fcc-Co and bismuth ferrite oxide (BFO). \label{tab:exchange} } \end{center} \end{table} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Spin orbit coupling}\label{subapp1:soc} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% The spin-orbit coupling is of fundamental importance when SD-MD simulations are at stake. This is, for example, the case with magnetocrystalline anisotropy, or the Dzyaloshinskii-Moriya interaction \cite{yosida_theory_1996}. Even if its non-relativistic origin corresponds to energies that are orders of magnitude lower than the exchange interaction, many technological application rely on effective interactions arising from that spin-orbit coupling \cite{skomski_simple_2008}. Depending on the crystalline lattice of the material under study, different forms of local magnetic anisotropies can arise. A first simple phenomenological model accounting for the one-site magnetocrystalline anisotropy consists of shaping the angular dependence of the corresponding energy surface with spherical harmonics. As an example, for uniaxial anisotropy, one has: \begin{equation} \mathcal{H}_{an}= -\sum_{{ i}=1}^{N} K_{an}(\bm{r}_{ i})\,\left(\bm{s}_{i} \cdot \bm{n}_{i} \right)^2, \end{equation} with $\bm{n}_{i}$ the direction of the anisotropy axis for the spin $i$, and $K_{an}(\bm{r}_{i})$ its anisotropic constant (in eV), which depends eventually to the position of the spin i. With this form of interaction, and depending on the sign of $K_{an}(\bm{r}_{i}) $, the result can be either an easy axis or an easy plane for the magnetization (easy axis if $K_{an}(\bm{r}_{i}) > 0$, easy plane for $K_{an}(\bm{r}_{i}) < 0$). However, in most magnetic crystals, the magnetocrystalline anisotropy takes more complex forms (like cubic anisotropy for example). Besides, this anisotropy model does not exhibit a clear dependence on the lattice parameters. Another, more sophisticated model aimed at taking into account the magnetocrystalline anisotropy is the two-site non-local Néel anisotropy \cite{neel_approche_1954}. Limited to the pseudo-dipolar term only, one has: \begin{eqnarray} \mathcal{H}_{N\acute{e}el}&=&-\sum_{{ i,j=1,i\neq j}}^Ng_1(r_{ij})\left(({\bm e}_{ij}\cdot{\bm s}_{i})({\bm e}_{ij}\cdot{\bm s}_{j})-\frac{1}{3}{\bm s}_{i}\cdot{\bm s}_{j}\right)\nonumber, \end{eqnarray} with $g_1(r)$ a fast decreasing function of $r$. Terms like $\sum_{{ i,j=1,i\neq j}}^Ng_1(r_{ij}){\bm s}_{i}\cdot{\bm s}_{j}$ can be viewed as special case of exchange interaction and can be omitted, by considering a well-suited redefinition of $J({r_{ij}})$. In our implementation, $g_1(r)$ was chosen to be the same function of three parameters as for the exchange interaction (see eq.~\ref{BSfunction}). It is also well known that the combination of the exchange interaction and the spin-orbit coupling can give rise to non-collinear spin states. First an object of theoretical work, this canted ferromagnetism (the ground state presents spins that are not perfectly aligned, but slightly tilted with one another) has since become a promising effect for many potential applications \cite{muhlbauer_skyrmion_2009,fert_skyrmions_2013}. The most common way to simulate this effect is to couple the exchange interaction (see Section.~\ref{subsec1:dynamics} and \ref{app:exchange}) to another interaction, referred to as the anti--symmetric Dzyaloshinskii-Moriya interaction \cite{dzyaloshinsky_thermodynamic_1958,moriya_anisotropic_1960}. This interaction takes the following form: \begin{equation} \mathcal{H}_{dm} = \sum_{{ i,j}=1,i\neq j}^{N} \bm{D}(r_{ij})\cdot\left(\bm{s}_{i}\times \bm{s}_{j}\right), \end{equation} with $\bm{D}(r_{ij})$ the Dzyaloshinskii-Moriya vector, which defines both the intensity (in eV) and the direction for the effect. In particular, the Dzyaloshinskii-Moriya interaction is known to be a key mechanism in the stabilization of magnetic skyrmions \cite{rohart_skyrmion_2013}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Magneto--electric interaction}\label{subapp1:me} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In some materials, like BFO, magnetism coexists and interplays with dielectric permanent polarization at the atomic scale. According to Katsura \emph{et al.} \cite{katsura_spin_2005} and Mostovoy \cite{mostovoy_ferroelectricity_2006}, the effects of this interplay can be taken into account within SD-MD via an anti--symmetric spin--spin effective interaction, as a particular case of the Dzyaloshinskii-Moriya vector, such as: \begin{equation} \mathcal{H}_{me} =\sum_{i,j=1,i\neq j}^{N} \left(\bm{E}\times\bm{e}_{ij} \right)\cdot\left(\bm{s}_{i}\times \bm{s}_{j} \right), \end{equation} with $\bm{E}$ giving the direction and the intensity of a screened dielectric atomic polarization (in eV). This polarization can also be induced by an external electric field of sufficient intensity. %%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Dipolar interaction}\label{subapp1:dipolar} %%%%%%%%%%%%%%%%%%%%%%%%%%%% With larger atomic systems, the long-range dipolar interaction is responsible for the stabilization of domain walls and the nucleation of magnetic domains. As such, it is often referred to as a demagnetizing field. Its general non-local expression is \begin{equation} \mathcal{H}_{di} = -\frac{\mu_0 \mu_B^2}{4\pi}\sum_{ i,j=1,i\neq j}^{N} \frac{g_{i} g_{j}}{\bm{r}_{ ij}^3} \left((\bm{s}_{i}\cdot\bm{e}_{ij} ) (\bm{s}_{ j}\cdot\bm{e}_{ij} ) -\frac{1}{3}(\bm{s}_{ i}\cdot\bm{s}_{i} ) \right), \end{equation} with $g_{i}$ and $g_{j}$ the Land\'e factors for spins$i$ and $j$ respectively, $\bm{r}_{ij}=\bm{r}_{j}-\bm{r}_{i}$, and $\bm{e}_{ij}=\bm{r}_{ij}/|\bm{r}_{ij}|$. Despite its simple formula, its computational cost is one of the main limiting factors for large magnetic simulations. Indeed, due to its long-range nature, the dipolar interaction does not have a finite cutoff distance, and therefore scales computationally as $N^2$, with $N$ the number of atoms in the system. To compute the dipolar effective field and force, two solutions have been considered. The first one takes advantage of the properties of the Suzuki--Trotter decomposition to avoid computing the dipolar interaction at each timestep. In MD simulations, this solution is often considered when thermostats or barostats have very different characteristic time scales, and is referred to as the r-RESPA algorithm \cite{tuckerman_reversible_1992}. A second one relies on an assumption of periodicity in space to use the well known technique of Ewald sums \cite{frenkel_understanding_2002}. For now, neither of these solutions are implemented for the SD-MD package in LAMMPS, but are being actively tested. Note that because the intensity of the dipolar interaction is usually much smaller than other considered magnetic interactions, in magnetic systems below the paramagnetic limit that are small enough to avoid the nucleation of domain walls, this effect can be safely omitted. %However, the intensity of the dipolar interaction is usually much smaller than the other magnetic interactions. %Its influence becomes non-negligible when the simulation of large magnetic systems is at stake (the size of a domain wall for example). %Thus, in most simulations, it is possible to neglect the influence of the dipolar long-range dipolar interaction. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Parametrization of a mechanical potential}\label{app:mechpotential} In MD, the standard approach to simulate properties of metals is to consider a mechanical interaction between the magnetic atoms using a empirial many-body potential such as the embedded atom method (EAM) \cite{daw_embedded-atom_1993}. Dudarev {\it et al} followed this strategy and developed a very specific semi-empirical many-body interatomic potential suitable for large scale molecular dynamics simulations of magnetic $\alpha$-iron \cite{dudarev_`magneticinteratomic_2005}. The functional form of the embedding portion of that potential is derived using a combination of the Stoner and the Ginzburg–Landau models, and reveals the spontaneous magnetization of atoms by a broken symmetry of the solutions of the Ginzburg–Landau model. Even if this strategy provides a link between magnetism and interatomic forces, the anisotropy effect through the spin-orbit coupling does not emerge naturally. This is not the strategy we have followed in this paper because we consider the magnitude of each atom's magnetic moment to be fixed during the simulation. Only its direction changes over time. However, there is no conceptual difficulty with implementing the more detailed potential in LAMMPS. Because EAM potentials are either fitted to experimental or {\it ab initio} data, the influence of the magnetic interactions is already silently included. In the framework of {\it ab initio} derived EAM potentials, this can be seen in the special case of collinear magnetism. Thus the standard approach is to effectively subtract the magnetic interactions from the mechanical potential, as~: \begin{equation} \mathcal{H}_{mech}^{eff}(\bm{r}_{i}) = \mathcal{H}_{mech}(\bm{r}_{i}) - \mathcal{H}_{mag}^{ground}(\bm{r}_{i}), \label{app:mech1} \end{equation} with $\mathcal{H}_{mech}^{eff}(\bm{r}_{i})$ the mechanical potential that needs computing, $\mathcal{H}_{mech}(\bm{r}_{ i})$ the EAM potential fitted before the magnetic energy subtraction, and $\mathcal{H}_{mag}^{ground}(\bm{r}_{i})$ the magnetic ground-state energy value. For example in a ferromagnet, because the exchange energy is by far the most intense value, the associated ground state is given by the following sum, with $j$ the neighboring atoms of the atom $i$: \begin{equation} \mathcal{H}_{mag}^{ground}(\bm{r}_{i}) = \sum_{{ j}=1}^{Neigh} {J}(r_{ij})\, |\bm{s}_{i}|\, |\bm{s}_{j}|. \end{equation} Future work will study how the combination of both experimental and \emph{ab initio} results can be used to derive better empirical magneto-mechanical potentials. \end{document} %% %% End of file `elsarticle-template-1-num.tex'. }
\caption{\textcolor[rgb]{0.00,0.00,0.00}{Compare our reconstruction result with ~\cite{li2009robust} in various of scene.}}
\caption{Illustration of verifying scores 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, e.g., 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 provided by verifier (showing in \textcolor{red}{red}).}
\caption{Verification-based detection. When an unreliable tracking result is found (showing in \textcolor{blue}{blue} in (a)), the verifier $\VF$ searches/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{Comparisons with pseudo real-time tracking methods on OTB2015 \cite{wu2015object} in distance precision rate (DPR\%) at a threshold of 20 pixels, overlap success rate (OSR\%) at an overlap threshold of 0.5, center location error (CLE) in pixels and speed (fps). The best two results are highlighted in \textcolor{red}{red} and \textcolor{blue}{blue} fonts, respectively}
\caption{Average DPR (\%) in terms of individual attributes on OTB2015~\cite{wu2015object}. The best two results are highlighted in \textcolor{red}{red} and \textcolor{blue}{blue} fonts, respectively}
\caption{Average OSR (\%) in terms of individual attributes on OTB2015~\cite{wu2015object}. The best two results are highlighted in \textcolor{red}{red} and \textcolor{blue}{blue} fonts, respectively.}
\caption{Comparisons with state-of-the-art tracking methods on VOT2016~\cite{kristan2016visual} in terms of expected average overlap (EAO\%), accuracy (\%), robustness (\%) and no-reset average overlap (AO\%). The best two results are highlighted in \textcolor{red}{red} and \textcolor{blue}{blue} fonts, respectively.}
\caption{{\bf Architecture.} \textcolor{myred}{Red}, \textcolor{myblue}{blue}, \textcolor{mygreen}{green} blocks indicate the non-shared portion of the CNN while the black ones are shared across all the streams for different datasets. Detailed specifications on the architecture are denoted on the last (\textcolor{mygreen}{green}) stream. {\bf B} and {\bf C} indicate the number of spectral bands and the number of classes, respectively, which are different for all the datasets.}
\caption{Monarch at measurement rate 4\%. (b) and (c) are of block-wise. (d) is of full-image. They all use MSE loss. (e) and (f) are improved with perceptual loss. They have stronger structure information than the state-of-the-art result in FCMN. Specially, we can see in the \textcolor[rgb]{1.00,0.00,0.00}{red circle} of (f), compared with (a) and (d), that even blurry image can be enhanced.}
\caption{Example for \leastsquares{} interpolation from smooth error \label{fig:lsinterpolation}}
\caption{The effect of penalization on a \leastsquares{} problem with solution $\widehat{p}$; (left) solution $q$ with $2$-norm penalization $\norm{q} < t$ (right) solution $p$ with $1$-norm penalization $\norm[1]{p} < t$. In the $2$-norm penalized solution we find $q_{1},q_{2}\neq 0$, but due to the shape of the $1$-norm unit cell we find $p_{2} = 0$ in the $1$-norm penalized solution.\label{fig:unitspheres}}
\caption[Solution of a two-dimensional \leastsquares{} problem by \LARS{}]{Solution of a two-dimensional \leastsquares{} problem by \LARS{}. In the first iteration \LARS{} proceeds along $w_{1}$ until the residual bisects the angle between $w_{1}$ and $w_{2}$. In the second and final iteration \LARS{} proceeds along the bisector.\label{fig:2dlar}\label{fig:largeometric}}
\caption{\label{fig:larexample} An example of the coefficient traces of a \LARS{} iteration with respect to accumulated $\ell_1$ coefficient changes along the \LARS{} iteration; % for six nearest neighbors and eight test vectors. (left) penalized \leastsquares{} coefficients (\LARS{} $\alpha$); (middle) unpenalized \leastsquares{} coefficients ($x + d_{\mathcal{A}}$ in \LARS); (right) correlations $\rho$.}
\caption{Many-particle wave functions of the excited states associated with the peaks in the optical absorption spectra of quasi-planar ($C_{3v}$) isomer of B$_{12}$ (Fig. 1a of the main article). $E$ denotes the excitation energy (in eV) of the state involved, and $f$ stands for oscillator strength corresponding to the electric-dipole transition to that state, from the ground state. In the ``Polarization'' column, x, y, and z denote the the direction of polarization of the absorbed light, with respect to the Cartesian coordinate axes defined in the figure below. In the ``Wave function'' column, \textbf{$|HF\rangle$} indicates the Hartree-Fock (HF) configuration, while other configurations are defined as virtual excitations with respect to it. $H$ and $L$ denote HOMO and LUMO orbitals, while the degenerate orbitals are denoted by additional subscripts 1, 2 etc. In parentheses next to each configuration, its coefficient in the CI expansion is indicated. Below, GS does not indicate a peak, rather it denotes the ground states wave function of the isomer. ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\protect\includegraphics[scale=0.12]{quasi_planar}}
\caption{Many particle wave functions of the excited states associated with the peaks in the optical absorption spectra of the double ring ($D_{6d}$) isomer of B$_{12}$ (Fig. 1b, main article). $E$ denotes the excitation energy (in eV) of the state involved, and $f$ stands for oscillator strength corresponding to the electric-dipole transition to that state, from the ground state. In the ``Polarization'' column, x, y, and z denote the the direction of polarization of the absorbed light, with respect to the Cartesian coordinate axes defined in the figure below. In the ``Wave function'' column, \textbf{$|HF\rangle$} indicates the Hartree-Fock (HF) configuration, while other configurations are defined as virtual excitations with respect to it. $H$ and $L$ denote HOMO and LUMO orbitals, while the degenerate orbitals are denoted by additional subscripts 1, 2 etc. In parentheses next to each configuration, its coefficient in the CI expansion is indicated. Below, GS does not indicate a peak, rather it denotes the ground states wave function of the isomer. ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\protect\includegraphics[scale=0.15]{double_ring}}
\caption{\label{fig:nozzle1and2mm_inj-fr}Ratio of flow rate injected into the main pipe through the concentric gap. The gap width is a) 1\,mm and b) 2\,mm. For increasing Reynolds numbers max and min depict the maximum and minimum injection flow rates that cause total relaminarization. All flow rate ratios in between also lead to relaminarization. For the cases c1-c4 at$\Rey=4000$ and $\Rey=5000$ in (b) marked by a {\color{red}+} see the text.}
\caption{\label{fig:inj_wall_2mm_profile_evo}Streamwise velocity profiles $(\bar w)$ measured at $z=2.5$ and (a) $\Rey=4000$ and (b) $\Rey=5000$. Cases $1-4$ depict different injection flow rates with the 2\,mm-FMD as indicated in figure\ref{fig:nozzle1and2mm_inj-fr} (b) by a {\color{red}+}. A measured profile of uncontrolled turbulent flow is shown for reference (ref).}
\caption{\label{fig:wall_inj_2mm_wrms_evo}Streamwise velocity fluctuations ($w_{rms}^2$) measured at $z=2.5$ and (a) $\Rey=4000$ and (b) $\Rey=5000$. Cases c1$-$c4 depict different injection flow rates as indicated in figure \ref{fig:nozzle1and2mm_inj-fr} (b) by a {\color{red}+}. The dashed line is showing uncontrolled turbulent flow for reference.}
\caption{\label{fig:k-re=5000}Local acceleration parameter $K$ for different injection flow rates by FMD\,2, measured at$\Rey=5000$ and $z=2.5$. Filled symbols indicate full relaminarization, half-filled symbols partial relaminarization further downstream.}
\caption{\label{fig:streakcontours}(Colour online) Contours of the (instantaneous) streamwise velocity $w$ in the cross-section of the pipe. The laminar flow field has been subtracted to emphasize near wall streaks (color bar expressed in units of the bulk velocity). a) is the unmodified turbulent reference flow and b) depicts case c\,2 with an injection flow rate as indicated in figure\ref{fig:nozzle1and2mm_inj-fr} (b) by a {\color{red}+} ($\Rey=4000$, $z=2.5$). $y^{+}\approx60$ is indicated by the dashed circle.}
\caption{\label{fig:transientgrowth}Variation of the transient growth with different injection flow rates, measured right behind (at $z=2.5$) the 2\,mm-FMD. Filled symbols indicate full relaminarization further downstream.}
\caption{Cumulative distribution function of various factors across \color{blue}{FWD} \color{black}and \color{red}{NFWD} \color{black}papers}
\caption{The Euclidean distance matrix of the 134 question groups where each group is represented by a word label. % Each row (column) represents the sampled questions that have the word label as the answer. % A matrix entry indicates the empirical expectation of the distance between the channel masks of the sentences from two question groups. % The labels are arranged into three topics: {\color{red} color}, {\color{myGreen} object}, and {\color{blue} spatial relation}. % A small distance indicates that the two channel masks are similar. % (Zoom in on the screen for a better view.) }
\caption{Comparing with existing methods on the 1st and 3rd category of 300VW benchmark. The \textcolor{cfirst}{1st}, \textcolor{csecond}{2nd} and the \textcolor{cthird}{3rd} place for each metric are color coded. Here AUC denotes the area under the CED curves in Fig \ref{fig:curves} and FR denotes the failure rate in percentage.}
\caption{Conditional prediction. Our algorithm fails on images with severe motion blur (a) \& (b). However, if an annotator gives hint of where the nose tip is (the yellow cross in (b)), we can correct the mistake (c) by finding the most probable class{\em conditioned} on this evidence.}
\caption{(Color online) \blue{(a) and (c) The distributions of the electron density (blue) in the $y=0$ plane at different propagation distances, superimposed with energetic electrons (color dots) with $\gamma>10$ and low energy electrons (black dots) with $\gamma<10$. (b) and (d) The distributions of the wakefield azimuthal magnetic field (normalized to $m_e\omega_p/e=4.237\times10^{2}$ T) in the $y=0$ plane at different propagation distances, the olive curves indicate the typical orbits of energetic electrons in the co-moving frame ($x, y, z-ct$). The imposed ETMF is $B_y^{\rm{ext}}=50$ T.} }
\caption{(Color online) \blue{The representative trajectories of two injected electrons in the $y=0$ plane are displayed, with the instantaneous (a) electric field in x direction and (b) magnetic field in y direction. The black circle denotes the turning points, the arrows denotes the directions of the magnetic force $F_\perp^B=ev_zB_y$ or electric force $F_\perp^E=-eE_x$.} (c) The minimum negative ($B_{y,\min}^{\rm{self}}$) and maximum positive ($B_{y,\max}^{\rm{self}}$) self-generated magnetic field in the focusing region vs the laser propagation distance. (d) $b_{\rm{crit}}$ predicted by Eq. (\ref{bmin}) in the cases $b_0$=0 and 0.117 ($B_y^{\rm{ext}}=50$ T). }
\caption{(Color online) \blue{(a) The trajectories of randomly selected 100 injected electrons with an ETMF $B_y^{ext}=20T$.} (b) The initial ionization positions of injected electrons in (a) (red stars), in comparison with those without the ETMF (black dots). (c) The trajectories of four typical injected electrons in (a) in the commoving frame. These four electrons originate from two ionization phases, respectively. }
\caption{Additional \aastex\symbols}
\caption{Additional \aastex\symbols}
\caption{{\color{blue}Hardness Intensity Diagram for the four outbursts during 2002-11. The spectra providing good fit with model M1, M2, M3, M4 and, M5 are shown with red, green, black, cyan, and pink respectively. The spectra with $\Gamma<1.8$ and $\sigma>1.5$ keV are shown with blue.}}
\caption{Instantaneous total energy density profiles, translational component (\sampleline{}), internal component (\sampleline{dashed}) and rotational (\textcolor{blue}{\sampleline{dotted}}), at (a) $1.8\delta$, (b) $3\delta$, (c) $4.2\delta$, and (d) $5.4\delta$ along the ramp}
\caption{Length of hypotheses through our NCPCM models versus absolute WER change.\\ Blue \& Green lines represent difference between WER of our system and the baseline ASR, for top-good and bottom-bad hypotheses, respectively. In an ideal scenario, all these lines would be below 0, thus all providing a change in WER towards improving the system. However we see in some cases that the WER increases, especially when the hypotheses length is short and when the performance is good. This is as expected since from Fig.~\ref{fig:domain_split} some cases are at 0\% WER due to the already highly-optimized nature of our ASR.\\ The red line represents the aggregate error over all data for each word length and as we can see in \textbf{all} cases the trend is one of \textbf{improving} the WER.%\\The dotted lines represent their respective trends. This matches {\color{red}Hypotheses D, E, F, G} \label{fig:wer_analysis}}
\caption{\label{tab:Classification-error-rates}Classification error rates of $k$-nearest neighbor classifier based on the metrics output by different methods. Three numbers below each datasets correspond to the feature dimensionality ($d$), number of classes ($c$) and number of examples ($n$). The best two results in each dataset are highlighted in {\color{red}red} and {\color{blue}blue}, respectively.}
\caption{Bugs from dataset Defects4J repaired by approaches built over Astor. In total, 98 bugs from 5 Java projects were repaired. \repairedsymbol{} means `bug with at least one test-suite adequate patch'. }
\caption{relative error of the potential of a charged conducting torus as computed by the toroidal and spherical series, \eqref{Vconducting toroidal}, \eqref{sph cond}. The blue circle is the edge of the torus and the points are \color{blue}$\circ$\color{black}: $\rho=R_0-r_0^2/R_0$, \color{blue}+\color{black}: $\rho=a$, \color{blue}$\ast$\color{black}: $\rho=R_0$. The toroidal series is computed up to $n=120$ and the spherical up to $k=170$ with $n=120$. The black regions are where both the spherical and toroidal harmonic solutions have converged to floating point accuracy. Increasing the number of terms would shrink the red and yellow regions, but numerical problems start to arise. We see the overlap of the regions of divergence in white - the inner torus and the spherical annulus. }
\caption{Steps in a typical MLC experiment. \highlight{1) Datasets have to be selected from the available repositories, taking into account their traits in order to choose the more appropriate ones for the task at hand. Some data preparation could be needed, depending on possible data peculiarities (i.e. an imbalanced dataset could be balanced through a resampling algorithm). 2) Methods related to the proposed one have to be selected and run, obtaining a set of predictions from each one of them. 3) These predictions have to be evaluated, picking from the large set of available performance metrics those more adequate to the analyzed problem. 4) It is advisable to draw conclusions including statistical tests over the previous results.}}
\caption{\highlight{General tips on how to avoid common mistakes during an experimentation in the multi-label field.}}
\caption{\textbf{Spatial correlations of transverse fluctuation.} The covariance $\text{cov}(\Delta t, \Delta x)$ of transverse fluctuations $r^k_{\perp, i}$ of beads at position $i$ on neighbouring chains $k$ and $l$ during a time interval $\Delta t$, as a function of the distance $\Delta x \,{=}\, \left| \mathbf{r}^k_{i}(t\,{+}\,\Delta t) - \mathbf{r}^l_{i}(t\,{+}\,\Delta t) \right|$, and normalised with respect to $\text{cov}(\Delta t, 0)$: $\text{cor}(\Delta t, \Delta x) \,{:=}\,\text{cov}(\Delta t, \Delta x)/\text{cov}(\Delta t, 0)$. Simulations were performed for an entangled solution with $\xi/L \,{=}\, 0.086$ and $\ell_p /L \,{=}\, 1$, and at different times $t$. As expected, for very small times $t \lesssim \tau_e$ (red curve), there is no correlation. In contrast, for times in the intermediate regime, $\tau_e \,{\lesssim}\, t\,{\lesssim}\,\tau_{eq}$ (here $t \,{=}\, 0.02\tau_d$), there are clear spatial correlations spanning a distance of about three times the mesh size $\xi$. These correlations vanish for times larger than the equilibration time, e.g.\ for times equal to the diffusion time, $t \,{=}\,\tau_d$. \label{fig:spatial_correlations}} \end{figure} % Our numerical data show clear spatial correlations in precisely that time window in which tube reorganisation takes place. Moreover, these correlations span several mesh sizes, strongly suggesting that many-body effects are responsible for tube reorganisation and the cocommittant relaxation of internal bending modes. We have also performed simulations for dilute solutions with $\xi /L \,{=}\, 0.51$ and found no correlations in the transverse displacements of neighbouring chains. Taken together, we conclude that constraint release (disentanglement) of internal modes is driven by the correlated motion of a given polymer and its surrounding polymers, and leads to an equilibration of internal bending modes on a time scale $\tau_{eq}$ smaller than the diffusion time $\tau_d$. As a consequence, for times larger than $\tau_{eq}$ the tube is dissolved insofar as the bending modes are equilibrated. However, this does not suffice for mechanical stresses to relax in a solution of semiflexible polymers, as the latter requires that correlations in the polymer's orientation also must vanish.\\ \textbf{Terminal relaxation.} To study this terminal relaxation regime we measured the auto-correlation function for the polymer orientation, $\delta{e}_R^2 (t) \,{:=}\,\langle \left[{\bf e} (t) {-} {\bf e} (0) \right]^2 \rangle$, where ${\bf e}(t) \,{=}\,{\bf R}(t) / R(t)$ denotes the instantaneous, normalised end-to-end vector of the polymer. Our simulations show that \textit{terminal relaxation} happens much later than the relaxation of the internal modes (Fig.~\ref{fig:terminal_relaxation}a). % \begin{figure}[t] \centering \includegraphics[width=\linewidth]{fig4_Lang_Frey.pdf} \caption{\textbf{Terminal relaxation.} {\bf a}, Time evolution of the mean-square end-to-end distance $\delta R^2 (t)$ (triangles), and the mean-square changes in the orientation $\delta e_R^2 (t)$ (circles) of a tracer polymer for entangled solutions with $\xi/L \,{=}\, 0.15$, $\ell_p /L \,{=}\, 5$ (red), and $\xi/L \,{=}\, 0.086$, $\ell_p /L \,{=}\, 0.44$ (yellow). Note that terminal relaxation sets in long after relaxation of internal bending modes. {\bf b}, Scaling plot of the terminal relaxation time $\tau_r / \tau_\text{Doi}$ with $\tau_\text{Doi} \,{=}\,\tau_d (L/\xi^4)$ as a function of $\ell_p/\ell_p^*$ with $\ell_p^* \,{=}\, L^3/\xi^2$, as determined from a fit of the asymptotic relaxation of $\delta e_R^2 (t)$ to the functional form $2{-}2 \exp(-t/\tau_r)$. The numerical data collapse to a master curve with two distinct regimes: For small values of $x \,{=}\,\ell_p/\ell_p^*$, the terminal relaxation time $\tau_r$ increases linearly with $x$, whereas for $ x \,{\gtrsim}\, 1$ it becomes independent of $x$, indicating that one has reached the rigid rod limit where Doi's scaling results are valid\cite{doi1978rods}. \label{fig:terminal_relaxation}} \end{figure} % To determine the terminal relaxation time $\tau_r$ we performed a least-squares fit of $\delta{e}_R^2 (t)$ close to saturation to $2{-}2 \exp(-t/\tau_r)$ (Supplementary Fig.~16), as expected for rotational diffusion\cite{granek1997semi}. For stiff chains, where the maximal mean-square amplitude of the bending modes $\delta{\bf r}_\perp^2 \,{\sim}\, L^3 /\ell_p$ is less than the squared mesh size $\xi^2$, we recover Doi's\cite{doi1978rods} result, $\tau_\text{Doi} \,{\sim}\,\tau_d (L/\xi)^4$ (Fig.~\ref{fig:terminal_relaxation}b). Hence, in this asymptotic limit, terminal relaxation is facilitated by Doi's \textit{`reptation -- tube rotation'} mechanism\cite{doi1978rods}: after diffusing its own length $L$ within a tube of diameter $d \,{\sim}\,\xi^2 / L$ in a time of the order of the diffusion time $\tau_d$, the rod becomes confined to a new tube, which is tilted relative to the previous one by an angle $\delta \theta \,{\sim}\, d/L$. The validity of Doi's effective reduction to a single-chain problem in the stiff limit has also been shown recently in computer simulations of needle-like rigid rods\cite{hofling2008entangled, Leitmann_etal:2016}. In contrast, for $\delta{\bf r}_\perp^2 \,{>}\,\xi^2$, where tube confinement of bending modes is significant, another relaxation mechanism sets in, which allows terminal relaxation to take place orders of magnitude faster. We find $\tau_r \,{\sim}\,\ell_p L^4 / \xi^2$ (Fig.~\ref{fig:terminal_relaxation}b, and Supplementary Fig.~17), which is clearly distinct from Odijk's\cite{odijk:83} result $\tau_\text{Odijk} \,{\sim}\,\ell_p L^2$ . Moreover, all our simulation data collapse on a universal scaling curve $\tau_r \,{=}\,\tau_\text{Doi} \,\hat \tau (\ell_p / \ell_p^*)$ (Fig.~\ref{fig:terminal_relaxation}b), confirming that there is a crossover from Doi's rigid rod scaling regime\cite{doi1978rods} to a qualitatively different \textit{disentanglement regime} by reducing the persistence length below some threshold value, $\ell_p \,{<}\,\ell_p^* \,{:=}\, L^3 /\xi^2$. This constitutes an intermediate asymptotic scaling regime extending over at least two decades ($10^{-3} \,{<}\, x\,{<}\, 1$) in the scaling variable $x = \ell_p / \ell_p^* = \ell_p \xi^2 / L^3$. It is precisely this regime which is most relevant for entangled solutions of cytoskeletal biopolymers\cite{hinner:98} as well as carbon nanotubes\cite{fakhri:10}. We can explain the observed terminal relaxation building on our results for the equilibration of internal modes by the following physical picture: After an initial phase of confinement to an Odijk-Semenov tube, the correlated motion of each polymer and its surrounding polymers leads to a reorganisation of the topological constraints, and thereby to an equilibration of internal bending modes, within a time $\tau_{eq} \,{\sim}\, L^5/\xi^2$. As a consequence, there is a mean-square angular rotation of the polymer due to bending fluctuations\cite{wilhelm1996radial, broedersz:14}, $\langle \delta \theta^2 \rangle \,{\sim}\, L/\ell_p$, and the ensuing rotational diffusion constant scales as $D_r \,{\sim}\,\langle \delta \theta^2 \rangle / \tau_{eq} \,{\sim}\,\xi^2 / (L^4 \ell_p)$. This is identical to the scaling of the terminal relaxation time as we find it in our simulations, $\tau_r \,{\sim}\,\ell_p L^4 / \xi^2$ (Fig.~\ref{fig:terminal_relaxation}b). Hence, terminal relaxation in solutions of entangled semiflexible polymers is not primarily due to diffusive motion along the tube's backbone (reptation), but rather due to \textit{disentanglement} of internal bending modes mediated by the release of topological constraints and the ensuing rotational diffusion of the polymer's orientation, induced by bending modes ($\langle \delta \theta^2 \rangle \,{\sim}\, L/\ell_p$). The constraint release for the internal modes is facilitated by the correlated dynamics of each polymer and its surrounding polymers. \\ \textbf{Comparison of entangled solutions with frozen environments.} In order to better understand the origin of these differences and the role of many-body effects in the dynamics of entangled solutions, we also studied the dynamics of a reference system where a tracer polymer moves in a frozen environment, similar as in recent experiments studying the Brownian motion of carbon nanotubes in a fixed agarose network\cite{fakhri:10}. In good quantitative agreement between our simulations in the frozen environment and the experimental results\cite{fakhri:10}, the rotational relaxation time exhibits a crossover from Doi scaling\cite{doi:78} to Odijk scaling\cite{odijk:83} with increasing polymer length $L$ (Fig.~\ref{fig:frozen}): % \begin{figure}[t] \centering \includegraphics[width=\linewidth]{fig5_Lang_Frey.pdf} \caption{\textbf{Scaling of the rotational relaxation time for a tracer polymer in an array of frozen polymer chains.} The numerical results for the rotational diffusion constant $D_r \,{\sim}\,\tau_r^{-1}$ in systems of different density and flexibility, rescaled as $(\tau_d/\tau_r) (\ell_p/L)$, are plotted as a function of rescaled polymer length $L/(\ell_p \xi^2)^{1/3} \,{=}\, (\ell_p^*/\ell_p)^{1/3}$. As in the experimental system, where carbon nanotubes diffuse in a porous agarose network (squares, data extracted from Ref.\cite{fakhri:10}), our simulations (solid circles) show that with increasing filament length $L$ there is a crossover from Doi's scaling result for rigid rods, $\tau_\text{Doi} \,{\sim}\, L^7/\xi^4$, (solid line) to Odijk's scaling result for semiflexible polymers, $\tau_\text{Odijk} \,{\sim}\,\tau_\text{Doi} \,\ell_p / L \,{\sim}\,\ell_p L^2$. To fit our definition, which uses the saturation time $\tau_r$ instead of the diffusivity, the experimental data has been rescaled by the necessary factor of $0.5$. Our simulations agree quantitatively well with the experimental data\cite{fakhri:10}. \label{fig:frozen}} \end{figure} % For rather stiff polymers with $ L/(\xi^2 \ell_p)^{1/3} \,{\lesssim}\, 2$, the relaxation time agrees with Doi's prediction, $\tau_r \,{\sim}\, L^7 /\xi^4$, while for semiflexible chains with $ L/(\xi^2 \ell_p)^{1/3} \,{\gtrsim}\, 2$, one finds Odijk's prediction, $\tau_r \,{\sim}\,\ell_p L^2$. Moreover, our Brownian dynamic simulations in a frozen environment are also consistent with previous numerical results of rigid rods in a random array of fixed obstacles where $\tau_r$ is reported to follow Doi's prediction\cite{munk:09, hofling2008enhanced, hofling2008entangled, nam2010reptation}. % \begin{figure}[!t] \centering \includegraphics[width=0.835\linewidth]{fig6_Lang_Frey.pdf} \caption{\textbf{Comparison between dynamics in a fixed array and an entangled solution}. Key quantities characterising the dynamics of a tracer polymer in a fixed array (green squares) or an entangled solution (yellow triangles) for a system with $\xi/L \,{=}\, 0.086$ and $\ell_p /L \,{=}\, 3$. To emulate a fixed array we used an equilibrated configuration of the entangled solution and froze the configuration of all but one tracer polymer, which is reminiscent of the experimental situation in Ref.\cite{fakhri:10}. In a fixed array, relaxation of a tracer polymer's orientation ($\delta e_R^2$) is much slower and follows the scaling predictions of Odijk and Doi. Moreover, the MSD of the polymer's fluctuations transverse to the end-to-end vector ($g_{1,\bot}$) exhibits a clearly visible plateau in a fixed array, but continues to relax in an entangled solution. At asymptotically large times, the transverse MSD and likewise the longitudinal MSD ($g_{1,\parallel}$) is slowed down in a fixed array. Finally, the relaxation of internal bending modes ($\delta R^2$) shows a qualitatively different behaviour: While there is an intermediate regime where $\delta R^2 (t)$ appears to saturate, which reflects caging in a tube, this is not the case for an entangled solution. There internal modes continuously relax, indicating ongoing tube reorganisation with concomitant relaxation of topological constraints. \label{fig:fixed_vs_solution}} \end{figure} % Comparing the dynamics in a fixed environment with the dynamics in an entangled solution, we find that \textit{all} key observables show qualitatively different behaviour (Fig.~\ref{fig:fixed_vs_solution}): The relaxation of the internal modes and the orientational correlations is retarded in a frozen environment, and the equilibration time $\tilde \tau_{eq}$, as well as the terminal relaxation time $\tilde \tau_r$, now agree well with the time it takes to diffuse the tube length, $\tilde \tau_{eq} \,{=}\,\tau_d$, and Odijk's scaling result, $\tau_\text{Odijk} \,{\sim}\,\ell_p L^2$, respectively. While for an entangled solution the correlation function for the transverse displacement ($g_{1,\perp}$) of the center of mass exhibits a slanted plateau, it shows a clear flat plateau in a frozen environment, indicating that the dynamics of the tracer polymer is confined to a persistent tube in the time window between the entanglement time $\tau_e$ and the tube diffusion time $\tau_d$. These predictions for the various key observables may be tested using microrheology experiments. Taken together, the comparison between the dynamics of entangled solutions and fixed arrays reaffirm that tube reorganisation, and the ensuing relaxation of topological constraints, must be based on the combined effect of the dynamics of a given polymer and its surrounding polymers. \section*{Discussion} The results presented here challenge our current understanding of the dynamics of entangled polymer solutions. While both asymptotic regimes, flexible polymers and rigid rods, are well described within the framework of classical reptation theory, our current data show that semiflexible polymer dynamics clearly exhibits strong many-body correlation effects. These correlation effects lead to a fast equilibration ('disentanglement') of the internal bending modes significantly before a polymer had time to diffuse the full length of its confining tube. As a consequence, terminal relaxation is facilitated by the correlated release of topological constraints and the ensuing rotational diffusion of the polymer orientation, and not by curvilinear motion as in classical reptation theory. The processes responsible for the restructuring and renewal of the tube appear to show similarities with dynamic correlations found in the glassy behaviour of dense colloidal systems\cite{berthier2011theoretical}. However, it remains a challenge to identify the essence of the many-body physics responsible for tube renewal leading to the relaxation of bending modes. To test the proposed disentanglement mechanism underlying terminal relaxation experimentally one actually does not need to measure the terminal relaxation time explicitly. It suffices to measure the equilibration time of the internal modes and validate the predicted scaling of the equilibration time. This should be feasible well within the time window accessible to experiments on entangled solutions of cytoskeletal filaments, carbon nanotubes, or short DNA filaments. \begin{acknowledgements} This project was supported by the German Excellence Initiative via the program ``NanoSystems Initiative Munich'' (NIM). \end{acknowledgements} \label{methods} \section*{Methods} \textbf{Brownian dynamics simulation.} We implemented a basic bead-spring algorithm\cite{fixman:78, grassia:95} to simulate the Brownian dynamics of polymer chains in an entangled solution, using standard interactions as discussed previously\cite{kremer1990dynamics, kremer1992simulations, paul1991dynamics, nam2010reptation}, and explained in detail below. Our simulations comprise $M$ polymer chains with $164 \,{\leq}\, M\,{\leq}\, 1106$ in a cubic simulation volume with edge length $X \,{=}\, 1.35\, L$, and periodic boundary conditions. Each polymer is represented by a linear sequence of $N$ beads connected by springs. For low densities we used $N \,{=}\, 45$, while for densities above $\xi/L \,{\leq}\, 0.1$ we used a finer discretisation with $N \,{=}\, 60$; this ensures that the polymers are thin enough such that the density of the entangled solution is sufficiently below the threshold to the nematic phase\cite{dijkstra1995simulation, khokhlov1981liquid, onsager:49}. In all our simulations we chose units such that $k_BT \,{=}\, 1$, and the friction per unit length is $\zeta \,{=}\, 1$. Also, we used the contour length of the polymers as our basic unit of length. For later reference, an actin filament with a contour length of $L \,{=}\, 10\,\mu \mathrm{m}$ and a diameter of $5 \,\mathrm{nm}$ in a solution with a viscosity of $\eta \,{=}\, 0.1\,\mathrm{Pa \, s}$ at a temperature of $20^\circ C$ has a disengagement time of $\tau_d \,{\approx}\, 1.5\times 10^4 \,\mathrm{s}$. The position of bead $i$ on polymer chain $k$ at time $t$ is denoted by $\mathbf{r}_{i}^{k}(t)$ with $1 \,{\leq}\, i\,{\leq}\, N$ and $1 \,{\leq}\, k\,{\leq}\, M$. Each bead $i$ is connected to its neighbours by FENE springs\cite{grest1986} with the interaction potential \begin{equation} \mathcal U^k_\text{FENE} = -B \sum_{1<i<N-1} \ln \left[ 1-\left( \frac{a^k_i-a_0}{a_\text{max}} \right)^2 \right] \, , \end{equation} where $a^k_i \,{=}\, |\mathbf{r}_{i+1}^{k} {-}\mathbf{r}_{i}^{k} |$ denotes the distance between two beads (bond length), with $a_0 \,{=}\, L/N$ the equilibrium bond length, and the maximum distance between beads set to $a_\text{max} \,{=}\, a_0/4$; for distances $|a^k_i| \,{>}\, a_0\,{+}\, a_\text{max}$ the FENE potential is $\mathcal U^k_\text{FENE} \,{=}\,\infty$. We chose the spring constant $B \,{\approx}\, 1000\, k_BT$. With this choice the deviation in bond length from the equilibrium value was always below $0.05 \, a_0$ in our simulations. The bending stiffness of the polymers is described by a standard worm-like chain model\cite{Kratky:1949, Saito:1967} with the bending energy given by \begin{equation} \mathcal U^k_{\mathrm{WLC}} = \frac{\ell_p k_BT}{a_0} \sum_{i=1}^{N-2} \left( 1 - \mathbf{t}^k_i \cdot \mathbf{t}^k_{i+1} \right) \, , \end{equation} where $\mathbf{t}^k_i \,{=}\, ({\mathbf{r}_{i+1}^{k} -\mathbf{r}_{i}^{k}})/{|\mathbf{r}_{i+1}^{k} -\mathbf{r}_{i}^{k}|}$ is a normalised bond vector (tangent vector). For the mutual (steric) interaction between the beads we used the Weeks-Chandler-Anderson (WCA) potential\cite{weeks1971role}, which for bead $i$ on chain $k$ reads for $r_{ij} \,{\leq}\,\sigma$ \begin{equation} \mathcal U^{i,k}_{\mathrm{WCA}} = A \sum\limits_{l,j} \left[ \left(\frac{\sigma}{r^{kl}_{ij}} \right)^{12} - 2 \left(\frac{\sigma}{r^{kl}_{ij}} \right)^{6} + 1 \right] \, , \end{equation} % $\sum\limits_{\substack{1<l<M \\ 1<j<N \\ j \neq i \ \mathrm{for}\ k=l} }$ where the sum extends over all other beads in the simulation box, i.e.\ over all polymers $l$ and beads $j$ except bead $i$ of chain $k$, and $r^{kl}_{ij} \,{=}\, |\mathbf{r}_{i}^{k} \,{-}\,\mathbf{r}_{j}^{l}|$ denotes the distance between a pair of beads $(i,j)$ on chains $k$ and $l$. For distances $r_{ij} \,{>}\,\sigma$, the potential vanishes: $\mathcal U^{i,k}_{\mathrm{WCA}} \,{=}\, 0$. We chose $\sigma \,{=}\, 0.9\, a_0$ for the range of the interaction. This choice prevents chain crossings and at the same time spurious oscillations of neighbouring beads in a chain due to their WCA interaction. In order to avoid crossing of chains or the overlapping of different beads we used a strong potential by setting the parameter $A \,{=}\, 20\, k_B T$ (see Supplementary Note 1). Finally, we implemented a cell-linked list algorithm to evaluate the occurring collisions which presorts the beads according to their positions before testing for polymer collisions. This led to a significant decrease in the runtime of our simulations. The Langevin equation for the entangled polymer solution reads \begin{eqnarray} \label{eq:eom} \zeta \, \frac{\partial \mathbf{r}^k_{i}(t)}{\partial t} = - \frac{\partial \, \mathcal{U}^{i,k}_{\mathrm{total}} ( \{\mathbf{r}^k_{i} \})}{\partial \mathbf{r}^k_{i}} + \boldsymbol{\eta}^k_i(t) \, , \label{eq:langevin} \end{eqnarray} where $\zeta$ denotes the friction coefficient of a single bead, $\boldsymbol{\eta}^k_i(t)$ is Gaussian white noise with mean zero and co-variances given by $ \langle \boldsymbol{\eta}^k_i(t) \,\boldsymbol{\eta}^l_j(t') \rangle = 6 \,\zeta\,\delta_{ij} \delta_{kl} \, k_B T\,\delta(t-t')$. The total potential acting on bead $i$ of polymer $k$ reads: $\mathcal{U}^{i,k}_{\mathrm{total}} \,{:=}\, U^{i,k}_{\mathrm{WCA}} \,{+}\,\mathcal U^k_\text{FENE} \,{+}\,\mathcal U^k_{\mathrm{WLC}}$. We used uniformly distributed random numbers generated by a maximally equidistributed combined Tausworthe generator\cite{galassi2011gsl} for the noise. These kinds of random number generators have been shown to amount to the same behaviour in the dynamics of polymers on time scales significantly above one time step as Gaussian white noise within the statistical errors while being significantly faster\cite{grassia:95, kremer1992simulations, ramanathan:07}. To calculate the time evolution, the Langevin equation, Eq.~\eqref{eq:langevin}, is integrated via a semi-implicit Euler algorithm\cite{euler} with a time step of (in our units) $1 \,{\times}\, 10^{-4}$ for systems with $N \,{=}\, 45$ and $2 \,{\times}\, 10^{-5}$ for $N \,{=}\, 60$, respectively. We performed extensive tests to ensure the reliability of our Brownian dynamics simulations (Supplementary Note 1). We find good agreement of both the tangent-tangent correlations (Supplementary Fig.~1) and the mean-square end-to-end distance (Supplementary Fig.~2) for freely relaxing polymers with known analytical results\cite{Kratky:1949, farge1993dynamic, frey1991dynamics, granek1997semi, hallat, kroy1997dynamic}. Moreover, we have tested that our simulation algorithm does not show spurious chain crossings (Supplementary Fig.~3). Finally, we have tested for finite size effects (Supplementary Fig.~4), and that our results for the mean-square displacement of the center monomer (Supplementary Fig.~5) and the mean-square changes of the end-to-end-vector are largely independent of the finite filament thickness (Supplementary Fig.~6). \textbf{Quantities of interest.} To characterise the dynamics of individual chains within the entangled polymer solution we studied the following quantities of interest where the averages $\langle \ldots \rangle$ indicate averages over all $M$ polymers in the simulation box, and over three independent realisations. We chose to characterise the dynamics of the internal bending modes of a tracer polymer in terms of its tangent-tangent correlation function, \begin{equation} C_{ij} := \langle \mathbf{t}_i \cdot \mathbf{t}_{j} \rangle \, , \end{equation} and the mean-square displacements (MSD) of the end-to-end distance $R^k = |{\bf R}^k|$, \begin{equation} \delta R^2 (t) := \langle \left[ R^k(t) -R^k(0) \right]^2 \rangle \, . \end{equation} In order to characterise the center-of-mass motion of a tracer filament we used the mean-square displacement of the center monomer parallel and perpendicular to the orientation of the end-to-end vector ${\bf e}^k \,{=}\,{\bf R}^k / R^k$, respectively: \begin{eqnarray} g_{1,\parallel}(t) & := &\left\langle \left[ \left(\mathbf{r}^k_{N/2}(t) -\mathbf{r}^k_{N/2}(0) \right) \cdot \mathbf{e}^k (0) \right]^2 \right\rangle \, , \\ g_{1,\bot}(t) & := & \left\langle \left[ \mathbf{r}^k_{N/2}(t) -\mathbf{r}^k_{N/2}(0) \right]^2 \right\rangle - g_{1,\parallel} (t) \, . \end{eqnarray} Finally, a quantity which allows to measure the terminal relaxation of stresses in the solution is given by the \textit{mean-square changes of the direction of the end-to-end vector} \begin{equation} \delta {e}_R^2 (t) \,{:=}\, \langle \left[ {\bf e}^k (t) {-} {\bf e}^k (0) \right]^2 \rangle \, . \label{eq:MSD_rot} \end{equation} \textbf{Definition of the covariance for transverse displacements.} To quantify the many-body effects in entangled polymer solutions, we were looking for correlations in the dynamics of neighbouring polymers. Since the tube is mainly constraining the fluctuations of a polymer transverse to its end-to-end vector, and the tube itself is due to the presence of neighbouring chains, we investigated correlations in these transverse fluctuations. We considered the magnitude of the displacement of bead $i$ on polymer $k$ during a time interval $\Delta t$, perpendicular to polymer's end-to-end vector $\mathbf{e}^k (t \,{+}\,\Delta t)$ \begin{equation} r^k_{\perp, i} := \left| \mathcal{P}_\perp^k (t+\Delta t) \cdot \left[ \mathbf{r}^k_{i}(t)- \mathbf{r}^k_{i}(t+\Delta t) \right] \right| \, , \label{eq:r_perp} \end{equation} where \begin{equation} \mathcal{P} _\perp^k (t) = 1-\mathbf{e}^k(t) \otimes \mathbf{e}^k (t) \end{equation} is a projection operator onto the end-to-end vector. Similar to the work of Doliwa and Heuer\cite{doliwa1998cage} on the cage effect in colloidal systems, we asked for correlations in the transverse displacements $r^k_{\perp, i}$ of neighbouring polymers, and define their covariance as %\cite{berthier2011theoretical, doliwa1998cage} \begin{equation} \text{cov}(\Delta t, \Delta x) := \left\langle r^k_{\perp, i} r^l_{\perp, i} \right\rangle - \langle r^k_{\perp, i} \rangle \langle r^l_{\perp, i} \rangle \, . \label{eq:def_orth_cov} \end{equation} The average in Eq.~\eqref{eq:def_orth_cov} is taken for a given chain $k$ with all other chains $k \,{\neq}\, l$ at a given time $\Delta t$ and a distance $\Delta x \,{=}\,\left|\mathbf{r}^k_{i}(t\,{+}\,\Delta t) - \mathbf{r}^l_{i}(t\,{+}\,\Delta t) \right|$, and we have also performed a moving time window average over a window of size $15 \, \tau_d$ discretised in subintervals of size $\Delta t$. Moreover, for simulations with $\Delta t \,{<}\, 0.2 \tau_d$ and $\Delta t \,{=}\, \tau_d$, we averaged over $4$ or $12$ independent realisations, respectively. For specificity, we used the beads at position $i \,{=}\, N/5$ and the equivalent beads at $i \,{=}\, 4N/5$. \begin{thebibliography}{10} \expandafter\ifx\csname url\endcsname\relax \def{\}{url}#1{\texttt{#1}}\fi \expandafter\ifx\csname urlprefix\endcsname\relax\def{\}{urlprefix}{URL }\fi \providecommand{\bibinfo}[2]{#2} \providecommand{\eprint}[2][]{\url{#2}} \bibitem{edwards:67} \bibinfo{author}{Edwards, S.~F.} \newblock \bibinfo{title}{The statistical mechanics of polymerized material}. \newblock{\bibinfo{journal}{Proc. Phys. Soc.}} \textbf{\bibinfo{volume}{92}}, \bibinfo{pages}{9} (\bibinfo{year}{1967}). \bibitem{degennes:71} \bibinfo{author}{de~Gennes, P.-G.} \newblock \bibinfo{title}{Reptation of a polymer chain in the presence of fixed obstacles}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{55}}, \bibinfo{pages}{572} (\bibinfo{year}{1971}). \bibitem{doi:78} \bibinfo{author}{Doi, M.} \&\bibinfo{author}{Edwards, S.~F.} \newblock \bibinfo{title}{Dynamics of concentrated polymer systems. {P}art 1. {B}rownian motion in the equilibrium state}. \newblock{\bibinfo{journal}{J. Chem. Soc. Faraday Trans.}} \textbf{\bibinfo{volume}{74}}, \bibinfo{pages}{1789--1801} (\bibinfo{year}{1978}). \bibitem{kremer1990dynamics} \bibinfo{author}{Kremer, K.} \&\bibinfo{author}{Grest, G.~S.} \newblock \bibinfo{title}{Dynamics of entangled linear polymer melts: A molecular-dynamics simulation}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{92}}, \bibinfo{pages}{5057} (\bibinfo{year}{1990}). \bibitem{mcleish:02} \bibinfo{author}{McLeish, T.} \newblock \bibinfo{title}{Tube theory of entangled polymer dynamics}. \newblock{\bibinfo{journal}{Adv. Phys.}} \textbf{\bibinfo{volume}{51}}, \bibinfo{pages}{1379--1527} (\bibinfo{year}{2002}). \bibitem{everaers:04} \bibinfo{author}{Everaers, R.} \textit{et~al.} \newblock \bibinfo{title}{Rheology and microscopic topology of entangled polymeric liquids}. \newblock{\bibinfo{journal}{Science}} \textbf{\bibinfo{volume}{303}}, \bibinfo{pages}{823--826} (\bibinfo{year}{2004}). \bibitem{hou2010stress} \bibinfo{author}{Hou, J.-X.}, \bibinfo{author}{Svaneborg, C.}, \bibinfo{author}{Everaers, R.} \&\bibinfo{author}{Grest, G.~S.} \newblock \bibinfo{title}{Stress relaxation in entangled polymer melts}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{105}}, \bibinfo{pages}{068301} (\bibinfo{year}{2010}). \bibitem{rubinstein:10} \bibinfo{author}{Rubinstein, M.} \newblock \bibinfo{title}{{Polymer physics -- The ugly duckling story: Will polymer physics ever become a part of {\textquotedblleft}proper{\textquotedblright} physics?}} \newblock{\bibinfo{journal}{J. Polym. Sci. B Polym. Phys.}} \textbf{\bibinfo{volume}{48}}, \bibinfo{pages}{2548} (\bibinfo{year}{2010}). \bibitem{broedersz:14} \bibinfo{author}{Broedersz, C.~P.} \&\bibinfo{author}{MacKintosh, F.~C.} \newblock \bibinfo{title}{Modeling semiflexible polymer networks}. \newblock{\bibinfo{journal}{Rev. Mod. Phys.}} \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{995} (\bibinfo{year}{2014}). \bibitem{odijk:83} \bibinfo{author}{Odijk, T.} \newblock \bibinfo{title}{The statistics and dynamics of confined or entangled stiff polymers}. \newblock{\bibinfo{journal}{Macromol.}} \textbf{\bibinfo{volume}{16}}, \bibinfo{pages}{1340--1344} (\bibinfo{year}{1983}). \bibitem{semenov:86} \bibinfo{author}{Semenov, A.~N.} \newblock \bibinfo{title}{Dynamics of concentrated solutions of rigid-chain polymers. {P}art 1. {B}rownian motion of persistent macromol. in isotropic solution}. \newblock{\bibinfo{journal}{J. Chem. Soc. Faraday Trans. 2}} \textbf{\bibinfo{volume}{82}}, \bibinfo{pages}{317--329} (\bibinfo{year}{1986}). \bibitem{romanowska:09} \bibinfo{author}{Romanowska, M.} \textit{et~al.} \newblock \bibinfo{title}{Direct observation of the tube model in {F}-actin solutions: Tube dimensions and curvatures}. \newblock{\bibinfo{journal}{Europhys. Lett.}} \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{26003} (\bibinfo{year}{2009}). \bibitem{ramanathan:07} \bibinfo{author}{Ramanathan, S.} \&\bibinfo{author}{Morse, D.~C.} \newblock \bibinfo{title}{Simulations of dynamics and viscoelasticity in highly entangled solutions of semiflexible rods}. \newblock{\bibinfo{journal}{Phys. Rev. E}} \textbf{\bibinfo{volume}{76}}, \bibinfo{pages}{010501} (\bibinfo{year}{2007}). \bibitem{granek1997semi} \bibinfo{author}{Granek, R.} \newblock \bibinfo{title}{From semi-flexible polymers to membranes: anomalous diffusion and reptation}. \newblock{\bibinfo{journal}{J. Phys. II}} \textbf{\bibinfo{volume}{7}}, \bibinfo{pages}{1761--1788} (\bibinfo{year}{1997}). \bibitem{fakhri:10} \bibinfo{author}{Fakhri, N.}, \bibinfo{author}{MacKintosh, F.~C.}, \bibinfo{author}{Lounis, B.}, \bibinfo{author}{Cognet, L.} \&\bibinfo{author}{Pasquali, M.} \newblock \bibinfo{title}{Brownian motion of stiff filaments in a crowded environment}. \newblock{\bibinfo{journal}{Science}} \textbf{\bibinfo{volume}{330}}, \bibinfo{pages}{1804--1807} (\bibinfo{year}{2010}). \bibitem{liu2006microrheology} \bibinfo{author}{Liu, J.} \textit{et~al.} \newblock \bibinfo{title}{Microrheology probes length scale dependent rheology}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{96}}, \bibinfo{pages}{118104} (\bibinfo{year}{2006}). \bibitem{semmrich:07} \bibinfo{author}{Semmrich, C.} \textit{et~al.} \newblock \bibinfo{title}{Glass transition and rheological redundancy in {F}-actin solutions}. \newblock{\bibinfo{journal}{Proc. Natl. Acad. Sci. U S A}} \textbf{\bibinfo{volume}{104}}, \bibinfo{pages}{20199--20203} (\bibinfo{year}{2007}). \bibitem{PerkinsChu1994Science} \bibinfo{author}{Perkins, T.~T.}, \bibinfo{author}{Quake, S.~R.}, \bibinfo{author}{Smith, D.~E.} \&\bibinfo{author}{Chu, S.} \newblock \bibinfo{title}{Relaxation of a single {DNA} molecule observed by optical microscopy}. \newblock{\bibinfo{journal}{Science}} \textbf{\bibinfo{volume}{264}}, \bibinfo{pages}{822--826} (\bibinfo{year}{1994}). \bibitem{hinner:98} \bibinfo{author}{Hinner, B.}, \bibinfo{author}{Tempel, M.}, \bibinfo{author}{Sackmann, E.}, \bibinfo{author}{Kroy, K.} \&\bibinfo{author}{Frey, E.} \newblock \bibinfo{title}{Entanglement, elasticity, and viscous relaxation of actin solutions}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{81}}, \bibinfo{pages}{2614} (\bibinfo{year}{1998}). \bibitem{kaes:94} \bibinfo{author}{K{\"a}s, J.}, \bibinfo{author}{Strey, H.} \&\bibinfo{author}{Sackmann, E.} \newblock \bibinfo{title}{Direct imaging of reptation for semiflexible actin filaments}. \newblock{\bibinfo{journal}{Nature}} \textbf{\bibinfo{volume}{368}}, \bibinfo{pages}{226--229} (\bibinfo{year}{1994}). \bibitem{perkins:94} \bibinfo{author}{Perkins, T.~T.}, \bibinfo{author}{Smith, D.~E.} \&\bibinfo{author}{Chu, S.} \newblock \bibinfo{title}{Direct observation of tube-like motion of a single polymer chain}. \newblock{\bibinfo{journal}{Science}} \textbf{\bibinfo{volume}{264}}, \bibinfo{pages}{819--822} (\bibinfo{year}{1994}). \bibitem{leGoffFrey2002prl} \bibinfo{author}{Le~Goff, L.}, \bibinfo{author}{Hallatschek, O.}, \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Amblard, F.} \newblock \bibinfo{title}{Tracer studies on {F}-actin fluctuations}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{89}}, \bibinfo{pages}{258101} (\bibinfo{year}{2002}). \bibitem{farge1993dynamic} \bibinfo{author}{Farge, E.} \&\bibinfo{author}{Maggs, A.~C.} \newblock \bibinfo{title}{Dynamic scattering from semiflexible polymers}. \newblock{\bibinfo{journal}{Macromol.}} \textbf{\bibinfo{volume}{26}}, \bibinfo{pages}{5041--5044} (\bibinfo{year}{1993}). \bibitem{frey1991dynamics} \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Nelson, D.~R.} \newblock \bibinfo{title}{Dynamics of flat membranes and flickering in red blood cells}. \newblock{\bibinfo{journal}{J. Phys. I (France)}} \textbf{\bibinfo{volume}{1}}, \bibinfo{pages}{1715--1757} (\bibinfo{year}{1991}). \bibitem{hallat} \bibinfo{author}{Hallatschek, O.}, \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Kroy, K.} \newblock \bibinfo{title}{{Propagation and relaxation of tension in stiff polymers}}. \newblock{\bibinfo{journal}{{Phys. Rev. Lett.}}} \textbf{\bibinfo{volume}{{94}}}, \bibinfo{pages}{77804} (\bibinfo{year}{{2005}}). \bibitem{kroy1997dynamic} \bibinfo{author}{Kroy, K.} \&\bibinfo{author}{Frey, E.} \newblock \bibinfo{title}{Dynamic scattering from solutions of semiflexible polymers}. \newblock{\bibinfo{journal}{Phys. Rev. E}} \textbf{\bibinfo{volume}{55}}, \bibinfo{pages}{3092} (\bibinfo{year}{1997}). \bibitem{wilhelm1996radial} \bibinfo{author}{Wilhelm, J.} \&\bibinfo{author}{Frey, E.} \newblock \bibinfo{title}{Radial distribution function of semiflexible polymers}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{77}}, \bibinfo{pages}{2581} (\bibinfo{year}{1996}). \bibitem{viovy1991constraint} \bibinfo{author}{Viovy, J.~L.}, \bibinfo{author}{Rubinstein, M.} \&\bibinfo{author}{Colby, R.~H.} \newblock \bibinfo{title}{Constraint release in polymer melts: {T}ube reorganization versus tube dilation}. \newblock{\bibinfo{journal}{Macromol.}} \textbf{\bibinfo{volume}{24}}, \bibinfo{pages}{3587--3596} (\bibinfo{year}{1991}). \bibitem{doi1978rods} \bibinfo{author}{Doi, M.} \&\bibinfo{author}{Edwards, S.~F.} \newblock \bibinfo{title}{Dynamics of rod-like macromolecules in concentrated solution. {P}art 1}. \newblock{\bibinfo{journal}{J. Chem. Soc. Faraday Trans.}} \textbf{\bibinfo{volume}{74}}, \bibinfo{pages}{560--570} (\bibinfo{year}{1978}). \bibitem{hofling2008entangled} \bibinfo{author}{H{\"o}fling, F.}, \bibinfo{author}{Munk, T.}, \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Franosch, T.} \newblock \bibinfo{title}{Entangled dynamics of a stiff polymer}. \newblock{\bibinfo{journal}{Phys. Rev. E}} \textbf{\bibinfo{volume}{77}}, \bibinfo{pages}{060904} (\bibinfo{year}{2008}). \bibitem{Leitmann_etal:2016} \bibinfo{author}{Leitmann, S.}, \bibinfo{author}{H{\"o}fling, F.} \&\bibinfo{author}{Franosch, T.} \newblock \bibinfo{title}{Tube concept for entangled stiff fibers predicts their dynamics in space and time}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{117}}, \bibinfo{pages}{097801} (\bibinfo{year}{2016}). \bibitem{munk:09} \bibinfo{author}{Munk, T.}, \bibinfo{author}{H{\"o}fling, F.}, \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Franosch, T.} \newblock \bibinfo{title}{Effective perrin theory for the anisotropic diffusion of a strongly hindered rod}. \newblock{\bibinfo{journal}{Europhys. Lett.}} \textbf{\bibinfo{volume}{85}}, \bibinfo{pages}{30003} (\bibinfo{year}{2009}). \bibitem{hofling2008enhanced} \bibinfo{author}{H{\"o}fling, F.}, \bibinfo{author}{Frey, E.} \&\bibinfo{author}{Franosch, T.} \newblock \bibinfo{title}{Enhanced diffusion of a needle in a planar array of point obstacles}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{101}}, \bibinfo{pages}{120605} (\bibinfo{year}{2008}). \bibitem{nam2010reptation} \bibinfo{author}{Nam, G.}, \bibinfo{author}{Johner, A.} \&\bibinfo{author}{Lee, N.-K.} \newblock \bibinfo{title}{Reptation of a semiflexible polymer through porous media}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{133}}, \bibinfo{pages}{044908} (\bibinfo{year}{2010}). \bibitem{berthier2011theoretical} \bibinfo{author}{Berthier, L.} \&\bibinfo{author}{Biroli, G.} \newblock \bibinfo{title}{Theoretical perspective on the glass transition and amorphous materials}. \newblock{\bibinfo{journal}{Rev. Mod. Phys.}} \textbf{\bibinfo{volume}{83}}, \bibinfo{pages}{587} (\bibinfo{year}{2011}). \bibitem{fixman:78} \bibinfo{author}{Fixman, M.} \newblock \bibinfo{title}{Simulation of polymer dynamics: {1} {G}eneral theory}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{1527--1537} (\bibinfo{year}{1978}). \bibitem{grassia:95} \bibinfo{author}{Grassia, P.~S.}, \bibinfo{author}{Hinch, E.~J.} \&\bibinfo{author}{Nitsche, L.~C.} \newblock \bibinfo{title}{Computer-simulations of brownian-motion of complex systems}. \newblock{\bibinfo{journal}{J. Fluid Mech.}} \textbf{\bibinfo{volume}{282}}, \bibinfo{pages}{373--403} (\bibinfo{year}{1995}). \bibitem{kremer1992simulations} \bibinfo{author}{Kremer, K.} \&\bibinfo{author}{Grest, G.~S.} \newblock \bibinfo{title}{Simulations for structural and dynamic properties of dense polymer systems}. \newblock{\bibinfo{journal}{J. Chem. Soc. Faraday Trans.}} \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{1707--1717} (\bibinfo{year}{1992}). \bibitem{paul1991dynamics} \bibinfo{author}{Paul, W.}, \bibinfo{author}{Binder, K.}, \bibinfo{author}{Heermann, D.~W.} \&\bibinfo{author}{Kremer, K.} \newblock \bibinfo{title}{Dynamics of polymer solutions and melts. reptation predictions and scaling of relaxation times}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{95}}, \bibinfo{pages}{7726--7740} (\bibinfo{year}{1991}). \bibitem{dijkstra1995simulation} \bibinfo{author}{Dijkstra, M.} \&\bibinfo{author}{Frenkel, D.} \newblock \bibinfo{title}{Simulation study of the isotropic-to-nematic transitions of semiflexible polymers}. \newblock{\bibinfo{journal}{Phys. Rev. E}} \textbf{\bibinfo{volume}{51}}, \bibinfo{pages}{5891} (\bibinfo{year}{1995}). \bibitem{khokhlov1981liquid} \bibinfo{author}{Khokhlov, A.~R.} \&\bibinfo{author}{Semenov, A.~N.} \newblock \bibinfo{title}{Liquid-crystalline ordering in the solution of long persistent chains}. \newblock{\bibinfo{journal}{Physica A}} \textbf{\bibinfo{volume}{108}}, \bibinfo{pages}{546--556} (\bibinfo{year}{1981}). \bibitem{onsager:49} \bibinfo{author}{Onsager, L.} \newblock \bibinfo{title}{The effects of shape on the interaction of colloidal particles}. \newblock{\bibinfo{journal}{Ann. N.Y. Acad. Sci.}} \textbf{\bibinfo{volume}{51}}, \bibinfo{pages}{627--659} (\bibinfo{year}{1949}). \bibitem{grest1986} \bibinfo{author}{Grest, G.~S.} \&\bibinfo{author}{Kremer, K.} \newblock \bibinfo{title}{Molecular dynamics simulation for polymers in the presence of a heat bath}. \newblock{\bibinfo{journal}{Phys. Rev. A}} \textbf{\bibinfo{volume}{33}}, \bibinfo{pages}{3628} (\bibinfo{year}{1986}). \bibitem{Kratky:1949} \bibinfo{author}{Kratky, O.} \&\bibinfo{author}{Porod, G.} \newblock \bibinfo{title}{R{\"o}ntgenuntersuchung gel{\"o}ster {F}adenmolek{\"u}le}. \newblock{\bibinfo{journal}{Recl. Trav. Chim.}} \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{1106--1122} (\bibinfo{year}{1949}). \bibitem{Saito:1967} \bibinfo{author}{Sait\^o, N.}, \bibinfo{author}{Takahashi, K.} \&\bibinfo{author}{Yunoki, Y.} \newblock \bibinfo{title}{Statistical mechanical theory of stiff chains}. \newblock{\bibinfo{journal}{J. Phys. Soc. Jpn.}} \textbf{\bibinfo{volume}{22}}, \bibinfo{pages}{219--226} (\bibinfo{year}{1967}). \bibitem{weeks1971role} \bibinfo{author}{Weeks, J.~D.}, \bibinfo{author}{Chandler, D.} \&\bibinfo{author}{Andersen, H.~C.} \newblock \bibinfo{title}{Role of repulsive forces in determining the equilibrium structure of simple liquids}. \newblock{\bibinfo{journal}{J. Chem. Phys.}} \textbf{\bibinfo{volume}{54}}, \bibinfo{pages}{5237} (\bibinfo{year}{1971}). \bibitem{galassi2011gsl} \bibinfo{author}{Galassi, M.} \textit{et~al.} \newblock \bibinfo{title}{{GSL}--{GNU} {S}cientific {L}ibrary: {R}eference manual} (\bibinfo{year}{2011}). \bibitem{euler} \bibinfo{author}{Giordano, N.~J.} \&\bibinfo{author}{Nakanishi, H.} \newblock{\bibinfo{title}{Computational Physics}} (\bibinfo{publisher}{Addison-Wesley, New York}, \bibinfo{year}{2005}). \bibitem{doliwa1998cage} \bibinfo{author}{Doliwa, B.} \&\bibinfo{author}{Heuer, A.} \newblock \bibinfo{title}{Cage effect, local anisotropies, and dynamic heterogeneities at the glass transition: A computer study of hard spheres}. \newblock{\bibinfo{journal}{Phys. Rev. Lett.}} \textbf{\bibinfo{volume}{80}}, \bibinfo{pages}{4915} (\bibinfo{year}{1998}). \end{thebibliography} \end{document} }
\caption{Label notation on the benchmark datasets \red{can be deleted}}
\caption{(a) Loss thresholds for a 4.8.8 lattice computed by checking percolation for the three logical operators (\protect\markerBlue{} \protect\markerRed{} \protect\markerGreen), branching~(\protect\markerTriangle) and the existence of solutions of Eq.~\eqref{eqn:ModifiedLogicalOp} (\protect\markerSquare) only for the red logical operator. Thresholds are plotted as a function of $1/d^{1/\nu}$, with $d$ the distance of the code and $\nu = 4/3$. The intercepts of the black lines (marked with the same symbols as the data) represent the critical threshold in the thermodynamic limit. (b) Phase diagram at $d = 36$ showing the probability to find a logical operator using the three methods explained in the main text. (c) Fraction of qubits left in the lattice for each of the three methods as a function of $1/d$. For methods (II) and (III), the fractions approach the fundamental 50\% limit.}
\caption{Top: Log$_{10}$ of the electron (solid) and ion (dashed) collision times in Myr as functions of radial position in \textcolor{red}{100} kpc for the profiles of $n$ and $T$ measured in A2199 by \citet{Johnstone2002}. Bottom: Log$_{10}$ of the collisionality parameter $k\lambda_i$ vs radial position in kpc, assuming a wavelength of 10kpc, for A2199. The damping rate of such a wave can be estimated accurately with fluid theory within about 100 kpc of the cluster, but should be calculated from kinetic theory beyond (see Figure \ref{figure6aab}) .}
\caption{\blue{Comparison of continuum mechanics results on normalized shear strain profile induced by a pre-sheared spherical inclusion (solid lines) with molecular dynamics data on spatial correlations of non-affine strain (open triangles). The CM data are plotted versus the normalized distance $x/a$ ($a$ is the inclusion radius) from the center of inclusion. The correlation function is plotted versus $r_\parallel/a_\text{eff}$, where $r_\parallel$ is the distance between two point within a parallel layer and $a_\text{eff}\approx 1.6\sigmaAA$ is the effective size of an elementary shear event (shear transformation zone). $z_\text{w}-z$ gives the distance from the wall. For clarity, part of the data is shifted upwards via multiplication by a factor of 4 (note the log-log scale). The analytic solution for the strain around an inclusion in an unbounded system~\cite{Dasgupta2013} confirms the validity of CM calculations away from the wall. As a further validation of CM calculations, MD results for the case of inclusion are also shown (filled symbols). The most striking result here is, however, the agreement between the behavior of the non-affine \emph{strain correlations} and the Eshelby-type strain field caused by an inclusion.}}
\caption{A comparison between the results of the exact formulas {\colorb{(solid lines)}} of $T(z)$ and $R(z)$ and their corresponding QNMs expansion (dashed lines) for below-resonance excitation $\omega_s < \omega_0$. The results are shown for different numbers of summation terms of the expansion.}
\caption{{\cred{Top)}} The field distribution of a Lorentz-dispersive resonator of parameters $\omega_0 L/c=10$, {\colorb{$\gamma L/c = 0.2$}}, $\Omega L/c=20$, and for a near-resonance harmonic excitation at frequency $\omega_s L / c=8$. The result of the QNMs expansion using 31 modes perfectly matches the exact formula. {\cred{Bottom)}} The field distribution of the five nearest % {\colorbb{\st{normalized} }} QNMs to the excitation frequency. }
\caption{ {\cred{ A comparison between the field distribution of QNMs formulation with the additional causality-related factor (solid lines) that shows no divergence outside the resonator, and the conventional QNMs formulation (dashed lines) that exhibits a divergence behavior. The results are shown for the two nearest QNMs to the excitation frequency of the example in Fig. 3. }}}
\caption[LoF]{Visual comparison of superpixels generated on Sintel images. The first row depicts the ground-truth optical flow, according to the gradient map \includegraphics[height=\baselineskip]{figs/gradient}, where the color and intensity represent, respectively, the orientation and magnitude. For each method named on the left board, the rows show the generated superpixels (top row) and the MUSE (bottom row), which is represented in heat map, according to the scale showed at the bottom of the Figure.}
\caption{Performance comparisons of DESPOT (\nscensymbol=100), GPU-DESPOT (\nscensymbol=1000), and HyP-DESPOT (\nscensymbol=1000) on the autonomous driving task. % \nscen represents the number of scenarios used in planning. }
\caption{\red{Example correlation surface} for the Barrier model by \cite{bakka2018}. The grey region acts as a physical barrier to spatial correlation, forcing the model to smooth around this barrier. }
\caption{The inset shows, as a function of the nanoparticles--surface distance $z$, the maximum of the ratio $G/G^{(0,0)}$ with respect to the interparticle distance $d$ for two SiC nanoparticles on top of a SiC substrate. In the main part of the figure, the same ratio is plotted as a function of $d$ for \textcolor{black}{$z=50\,$nm} (black solid line), 500\,nm (red dashed line),$1\,\mu$m (blue dotted line) and $10\,\mu$m (orange dot-dashed line).}
\caption{The inset shows, as a function of the nanoparticles--surface distance $z$, the maximum of the ratio $G/G^{(0,0)}$ with respect to the interparticle distance $d$ for two gold nanoparticles on top of a SiC substrate. In the main part of the figure, the same ratio is plotted as a function of $d$ for \textcolor{black}{$z=50\,$nm} (black solid line), 500\,nm (red dashed line),$1\,\mu$m (blue dotted line) and $10\,\mu$m (orange dot-dashed line).}
\caption{Panel (a) shows the conductance ratio $G/G^{(0,0)}$ as a function of $d$ between two SiC nanoparticles placed at distance \textcolor{black}{$z=50\,$nm} from a SiC substrate. The four lines correspond to the absence of graphene (black solid line), and to configurations with graphene having $\mu=0.1\,$eV (red dashed line), 0.3\,eV (blue dot-dashed line) and 0.5\,eV (orange dotted line). The inset shows the spectral conductance associated with the four same configurations. Panel (b) and its inset show the same quantities for two gold nanoparticles placed at distance\textcolor{black}{$z=50\,$nm} from a SiC substrate.}
\caption{A-F) Measured and fitted through-focus PSF of beads emitting in the green (range: [-1, 1] $\mu$m, 21 steps) before and after aberration correction, see also \textcolor{blue}{Visualisation~1} and \textcolor{blue}{Visualisation~2}. The left columns of A and D show measured and fitted focal slices, the right columns of A and D show the same images contrast stretched with the same scale for each exp/fit pair for visibility. Scalebar indicates 1~$\mu$m. G) Fitted Zernike coefficients (rms values) before and after aberration correction showing a reduction in the $W_\mathrm{rms}$ of the model PSFs that best fit the measurements from $59\pm 1$~m$\lambda$ to $13.4\pm 0.4$~m$\lambda$.}
\caption{\textcolor{comm}{Our measure of structural diradical character is based on comparing bond-length patterns of molecular structures with idealized bond-length patterns for closed- (top left) and open-shell diradical (top right) forms, shown here for p-quinodimethane. %Note that strictly speaking, these are not mesomeric forms, since bond lengths will differ between the diradicaloid and the closed-shell structures.}} }}
\caption{Relative energies with respect to the closed-shell energy ($\Delta E$) [kJ/mol], structural \textcolor{comm2}{diradical} character $y_{s}$ (Equation~\ref{eq:osc}) (MAE and, in parentheses, RMSD), bond-length alternation (Equation~\ref{eq:bla}) $\mathrm{BLA}$ [pm], $\hat{S}^{2}$ expectation values ($\braket{\hat{S}^{2}}$) for the optimized structures of \textbf{[Co\textsubscript{2}]} (see Figure~\ref{fig:switch}) employing BP86/def2-TZVP and B3LYP/def2-TZVP, for closed-shell (cs), open-shell singlet modeled by a broken-symmetry (bs) determinant, and triplet (t). The structural diradical characters of the energetically most stable structures are highlighted in green (with energies differing by less than 5 kJ/mol considered as degenerate). The overall assignment as closed-shell (CS) or open-shell (OS) is indicated in the right-most column.}
\caption{Quantitative comparisons with 15 methods on 4 large-scale 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. Our method (VGG-16) ranks first or second on these datasets. ``--'' means corresponding methods are trained on that dataset.}
\caption{\label{tab:timings}Runtime of the individual stages of our method for a single frame on Intel\textregistered Core\textsuperscript{\tiny TM} i7-5930K, 32GB, NVIDIA Titan X. For depth and motion estimation inference times only are indicated.}
\caption{Same as Fig.~\ref{fig:QNM_beta-5}, except for $\beta=100$. Panels (a) and (b) refer to Model 1 alone since all scalarized stars in Model 2 are unstable (see SM). For the sake of presentation, in panel (c), we plot the absolute value of the $\phi$-mode frequencies only, scaled by the sign of the imaginary part of $\omega$ to help distinguish between stable and unstable modes. Note that scalarized solutions in M2 with $n=0$ cease to exist at the same critical compactness $M/R \approx 0.292$, where scalarized solutions in M1 first appear (cf. Fig.~\ref{fig:background}\red{b}).}
\caption{Same as in Fig.~2, but for the square lattice. Here, however, the degeneracy of the ice-rule vertices ($q=0$) is lifted by a difference in interaction strength between perpendicular and collinear traps or dumbbells (green dotted lines), and polarized vertices have higher energy (middle: forth from left; bottom: second from left). The red \textcolor{red}{$\wedge$} connects two plaquettes where the head-to-toe rule is broken by a monopole (see Fig.~5).}
\caption{Top (a-c): Decimating a square lattice (a) {\it and} its anfiferromagnetic ground state (decimated traps are replaced with negatively saturated traps, in red) leads to an ordered lowest energy state with $\tilde q=-2$ virtual charges on half of the decimated vertices in the SI picture (b). It corresponds to $q=\pm 1$ real charges on decimated vertices in the PI picture (c) and thus all vertices obey the ice rule. At low decimation, this is the only low energy state. Bottom (d-f): However, above the decimation threshold corresponding to the percolation of decimated neighboring square plaquettes [e.g. yellow shaded ones neighboring a green one in (a)], the low energy state becomes degenerate. A disordered state can be chosen by connecting (red dotted line) neighboring decimated plaquettes (d) with $q=-2$ monopoles (represented with \textcolor{red}{$\wedge$} connectors as in Fig.~3) on $z=4$ vertices, thus removing the virtual charges from decimated, $z=3$ vertices without increasing the energy [(e), green rectangles frame spins that can be freely flipped]. In the PI picture (f) this corresponds to ice rule violations on the $z=4$ vertices hosting charge $q=2$: disorder comes from entropic transfer of topological charge.}
\caption{ %\footnotesize The structure of each Harmonious Attention module consists of {\bf (a)} Soft Attention which includes {\bf (b)} Spatial Attention (pixel-wise) and {\bf (c)} Channel Attention (scale-wise), and {\bf (d)} Hard Regional Attention (part-wise). Layer type is indicated by background colour: \textcolor{battleshipgrey}{\bf grey} for \textit{convolutional} (conv), \textcolor{brown}{\bf brown} for \textit{global average pooling}, and \textcolor{capri}{\bf blue} for \textit{fully-connected} layers. The three items in the bracket of a conv layer are: filter number, filter shape, and stride. % The ReLU \cite{krizhevsky2012imagenet} and Batch Normalisation (BN) \cite{ioffe2015batch} (applied to each conv layer) are not shown for brevity. }
\caption{Comparison of the measured \lal PDFs at $4.9<z<6.1$ with fully post-processed outputs of numerical simulations from \red{\citet{Chardin17}} (\textit{top}) and \citet{Keating17} (\textit{bottom}). The coloured contours show the envelope of the pessimistic and optimistic bounds for 100 realisations. Post-processing consists of randomly drawing a number of lines of sight from the simulations equal to the number of observations in the corresponding redshift range. The simulated lines of sight are then forward-modelled to mimic the observed spectra (see text).}
\caption{Posterior distributions on the skewness $s$ and mean opacity $\overline{\tau} = -\text{log} \ \overline{F}$ of \lal transmission. Different contours correspond to redshift ranges of $\Delta z = 0.2$ beginning at $z=4.9,5.1,5.3,5.5,5.7,5.9$, following the direction indicated by the arrow. The colored thick lines correspond to simulations from \red{\citet{Chardin17}} (\textit{orange}), \citet{Keating17} (\textit{red}) and \citet{Bolton17} (\textit{blue}), post-processed to mimic observational data as described in Section 5.1.}
\caption{Values of the scalar resistance functions over non-dimensional gap, $\xi$, for size ratio $\lambda=1$. Those generated from the near-field formulae are represented by hollow circles ({\color{blue}\Circle}), and those generated from Lamb's solution \citep{wilson_stokes_2013} are filled circles ({\color{blue}\CIRCLE}). The dashed vertical line appears at $\xi=\lambda$, recalling that the near-field formulae are only valid for $\xi \ll \lambda$.}
\caption{Values of the scalar resistance functions over non-dimensional gap, $\xi$, for size ratio $\lambda=0.1$. Those generated from the near-field formulae are represented by hollow circles ({\color{blue}\Circle}), and those generated from Lamb's solution \citep{wilson_stokes_2013} are filled circles ({\color{blue}\CIRCLE}). The dashed vertical line appears at $\xi=\lambda$, recalling that the near-field formulae are only valid for $\xi \ll \lambda$.}
\caption{Values of the scalar resistance functions over non-dimensional gap, $\xi$, for size ratio $\lambda=0.01$. Those generated from the near-field formulae are represented by hollow circles ({\color{blue}\Circle}), and those generated from Lamb's solution \citep{wilson_stokes_2013} are filled circles ({\color{blue}\CIRCLE}). The dashed vertical line appears at $\xi=\lambda$, recalling that the near-field formulae are only valid for $\xi \ll \lambda$.}
\caption{\label{fig: bif}(a) Bifurcation diagram showing the spatially averaged energy $\E$ of primary equilibria $E_n$ (\solidrule) and odd secondary equilibria ({\color{blue}\dashedrule}) of the KSE. Stability of equilibria is not indicated. (b-d) Steady solutions $u(x)$ on the $E_1$ branch at the three points indicated ({\footnotesize\color{red}$\bullet$}) in panel (a), where the values of $L$ are (b) 1.03, (c) 1.19, and (d) 1.70.}
\caption{\label{fig: trunc all} Upper bounds on the mean energy of odd solutions, computed for large truncations of the KSE using auxiliary functionals of degree 2 ({\color{blue}\tiny $\mathbf{\square}$}), 4 ($\circ$), and 6 ({\color{red}$*$}). Mean energies of primary equilibria are shown also ({\color{gray}\solidrule}). Degree-6 bounds are shown only in panel (b), which is a detailed view of the boxed region in panel (a).}
\caption{\label{fig: exponents} Local exponents of power laws $c L^p$ fit to upper bounds on mean energy for auxiliary functionals of degree 2 ({\color{blue}\tiny $\mathbf{\square}$}) and 4 ($\circ$). The bound values are shown in figure \ref{fig: trunc all}(a) over a different range of $L$.}
\caption{\label{fig: N}Upper bounds on mean energy computed using quartic $V$: (a) bounds $B_{N,4}^{pde}$ for the full KSE and (b) bounds $B_{N,4}$ for truncations of the KSE. The number of modes ($N$) used to compute the bounds are 4 ($\vartriangle$), 6 ({\color{cyan}$\triangledown$}), 8 ({\color{red}\large $\triangleleft$}), and 10 ({\color{blue}\large $\triangleright$}). Mean energies of primary equilibria are shown also ({\color{gray}\solidrule}).}
\caption{\label{fig: pde}Upper bounds on $\ov\E$ for the KSE computed by solving \eqref{eq: B pde kse} with quartic $V$. The bounds saturated by each of the first four branches of $E_n$ equilibria ({\color{gray}\solidrule}) were computed using $N = 6n$ modes.}
\caption{\label{fig: background} Optimal coefficients $\alpha$ and functions $\zeta(x)$ for the background method \eqref{eq: V opt quad} with various $L$, computed by solving the truncated optimization \eqref{eq: B trunc kse} with $2d=2$ and $N\ge8L$ modes. Displayed quantities are (a) optimal $\alpha$ for various domain sizes $2\pi L$, (b) optimal $\zeta(x)$ for $L=10^{1.3}$, and (c-d) details of optimal $\zeta(x)$ for $L=10^{1.3}$ (\solidrule), $10^{1.6}$ (\longdashedrule), and $10^{1.9}$ (\dottedrule).}
\caption{(a) Expectation ($\frac{~~~}{~~~}$) and one-standard-deviation bounds ($\frac{~}{~}\frac{~}{~}\frac{~}{~}$) computed from $p(y_{t+1}|y_t)$ for $\delta y = 0.04$ and $\lambda = 0.1$ and (b) the probability distribution estimated with $\lambda = 0$ ($\circ$) and $\lambda=0.1$ ({\color{red}$\bullet$}).}
\caption{(a) Next-step prediction ($\frac{~~~}{~~~}$) and 95\% confidence interval ($\frac{~}{~}\frac{~}{~}\frac{~}{~}$) from DE-LSTM. The solid circles denote the ground truth, $y(t)$. (b) DE-LSTM prediction ($\circ$) from the noisy data ({\color{red}$\bullet$}). The dashed line is the ground truth, $y(t)$. DE-LSTM is trained for $\delta y = 0.03 sd[\yhat]$ and $\lambda = 0.01$. }
\caption{(a) Sample trajectory of the forced Van der Pol oscillator and (b) auto-correlation functions; $\frac{~~~~~}{~~~~~}$, $\rho(\tau;y,y)$; {\color{blue}$\frac{\,~}{\,~}\,\frac{\,}{\,}\,\frac{\,~}{\,~}$}, $\rho(\tau;u,u)$; {\color{red}$\frac{~\,}{~}\,\frac{~\,}{~}\,\frac{~\,}{~}$}, $\rho(\tau;u,y)$.}
\caption{ Next-step prediction ($\frac{~~~}{~~~}$) and 95\% confidence interval ($\frac{~}{~}\frac{~}{~}\frac{~}{~}$) from DE-LSTM. The solid circles ({\color{red}$\bullet$}) denote the ground truth, $y(t)$, and the hollow circles ($\color{blue}\circ$) are the noisy observation, $\yhat_t$. }
\caption{\coloronline Wigner function of an angular momentum measurement in z- and x direction on a spin 3/2-system with $\psi= |m,j\rangle=|-1/2,3/2\rangle$. Gaussian regularization with $\veps=0.001$. Lower left: Level sets encoded with blue$\to$negative, red$\to$positive. Lower right: Marginal distributions in $z$-direction (red) and $x$-direction (blue).}
\caption{\coloronline Wigner function of a pair of two random $4\times4$-matrices with $\rho=\idty/4$. }
\caption{\label{fig:penta}\coloronline Wigner function of multiplet with 5-fold rotation and reflection symmetry, illustrating property 7. Further details in \suppl8.}
\caption{\label{fig:penta2}\coloronline The same example as Fig.~\ref{fig:penta} as an illustration of property~6, showing the connecting ellipses (thin black) and the exact singular manifold (light blue). }
\caption{\label{fig:spin4}\coloronline Wigner function of angular momentum components $L_x$ and $L_y$ for spin $s=4$, depicted for the maximally mixed state, the $j_x=0$ and the $j_x=4$ eigenstate. The Wigner fuction of the maximally mixed state shares the full rotational symmetry. The Wigner function of the $j_x=0$ state has only reflection symmetries on the $x=0$ and the $y=0$ axis. The Wigner function of the $j_x=4$ state is clearly concentrated in the $\langle L_x \rangle \geq 0$ region and has only a symmetry on the $y=0$ axis.}
\caption{\coloronline Left: Wigner function of two $3\times3$-matrices \eqref{heartMatrix}. Right: The singularity curve $\sing$, the quartic \eqref{heartquart}. Grey: the range $\jnr$. }
\caption{\coloronline Wigner function (top) and singular set (bottom) of a pair of matrices illustrating \Qref{Pty:nc}. The singular set $\sing$ is represented in thick light blue. Thin lines are the ellipses described in \Qref{Pty:nc}. Bottom right: blow-up of the upper right corner indicated in the left diagram. }
\caption{\label{fig:penta}\coloronline Wigner function of multiplet with 5-fold rotation and reflection symmetry, illustrating \Qref{Pty:cov}.}
\caption{Top 36 Association Rules for GitHub-specific vs. \colorbox{gray!25}{npm-specific packages}.}
\caption{% R vs T (and figures of merit), dR/dT and dip/peak temperatures Temperature dependence of the electrical transport. (a) $R_s$, normalized at $T=300$ K, as a function of $T$, before and after the onset of Li\apex{+} intercalation. (b) Overdrive voltage dependence of the $R_s$ at $T=300$ K (violet dots and lines), and of the residual resistivity ratio (green diamonds). Dashed line acts as a guide to the eye. (c) $T$ dependence of the first derivative of the $R_s$, before and after Li\apex{+} intercalation. {\color{blue}Curves are color-coded to match the legend in panel (a).} Arrows indicate the $T$ values where the first derivative shows a dip (red) and a peak (blue). Curves are shifted for clarity. (d) Overdrive voltage dependence of $T_{hump}$ (blue up triangles) and $T_{dip}$ (down red triangles) in (c).}
\caption{Greedy Design $\left( {{\boldsymbol{\theta }}, {T_A},\epsilon } \right)$} \label{Algo_1} \begin{algorithmic}[1] \renewcommand{\algorithmicrequire}{\textbf{Input:}}\REQUIRE ${{\bf{x}}}$, ${\boldsymbol{\theta }}$, ${T_A}$, $\epsilon$ \STATE Compute ${c^{{\text{LL}}}}$ from \eqref{c_LL} \STATE Initialize: ${{\bf{x}}^{{\text{adv}}}}\left(0 \right)\leftarrow{\bf{x}}$ % ---------------------------------------------------------------------- \FOR {$i = 1$ \TO $\left\lfloor{\epsilon{N_X}T} \right\rfloor$} { \STATE ${{\bf{x}}^{{\text{adv}}}}\left(i \right)\leftarrow{{\bf{x}}^{{\text{adv}}}}\left({i - 1} \right)+ {\bf{p}}$, where ${\bf{p}}$ is obtained by solving problem \eqref{adv_attack} with ${{\bf{x}}^{{\rm{adv}}}}\left({i - 1} \right)$ in lieu of ${\bf{x}}$ and $p_{j, t} = 0$ for all $t > T_A$. % ------------------------------------------------------------------ }
\caption{Adversarial Training $\left( {{T_A},{\epsilon_A}} \right)$} \label{Algo_2} \begin{algorithmic}[1] % ---------------------------------------------------------------------- \renewcommand{\algorithmicrequire}{\textbf{Input:}}\REQUIRE Training set, basis functions ${\bf{A}}$ and ${\bf{B}}$, learning rate ${\eta}$, ${T_A}$, and ${\epsilon _A}$ % ---------------------------------------------------------------------- \renewcommand{\algorithmicrequire}{\textbf{Initialize}:} \REQUIRE $\boldsymbol{\theta}$ % ---------------------------------------------------------------------- \FOR {\text{each iteration}} { % ------------------------------------------------------------------ \STATE Choose example $\left({{\bf{x}},{c}} \right)$ from the training set \STATE Compute ${\bf{x}}^{{\text{adv}}}$ and ${c^{{\text{LL}}}}$ from Algorithm~\ref{Algo_1} with input $\boldsymbol{\theta}$, ${T_A}$ and ${\epsilon _A}$ \STATE Update $\boldsymbol{\theta}$: $\boldsymbol{\theta} \leftarrow \boldsymbol{\theta} + \eta {\nabla _{\boldsymbol{\theta }}}{L_{\boldsymbol{\theta }}}\left( {{\bf{x}}^{{\rm{adv}}},{c}} \right)$ % ------------------------------------------------------------------ }\ENDFOR % ---------------------------------------------------------------------- \renewcommand{\algorithmicrequire}{\textbf{Output:}}\REQUIRE $\boldsymbol{\theta}$ % ---------------------------------------------------------------------- \end{algorithmic} \end{algorithm} % ------------------------------------------------------------------------------ % % % ============================================================================== \section {Robust Training}\label{sec:Robust_training} % ============================================================================== % In order to increase the robustness of the trained SNN to adversarial examples, in this section, we propose a robust training procedure. Accordingly, in a manner similar to \cite{madry2017towards}, during the SGD-based training phase, each training example $\left({{\bf{x}},{c}} \right)$ is substituted with the adversarial example ${\bf{x}}^{{\text{adv}}}$ obtained from Algorithm~\ref{Algo_1} for the current iterate $\boldsymbol{\theta}$. The training algorithm is detailed in Algorithm~\ref{Algo_2}. % Note that, the robust training algorithm is parameterized by $T_A$ and $\epsilon_A$, which determine the parameters of the assumed adversary during training. % % ============================================================================== \section{Numerical Results}\label{sec:Sim_sec} % ============================================================================== % In this section, we numerically study the performance of the described probabilistic SNN under the adversarial attacks. We use the standard USPS dataset as the input data. As a result, we have $N_X = 256$, with one input neuron per pixel of the $16 \times 16$ images. Unless stated otherwise, we focus solely in the classes $\{1, 5, 7, 9 \}$ and we set $T = K = 16$. We assume the worst-case $T_A = T$ for the adversary during the test phase. For rate decoding, we use a desired spike train with one spike after every three zeros. SGD is applied for 200 training epochs and early stopping is used for all schemes. Holdout validation with 20$\%$ of training samples is applied to select between $10^{-3}$ and $10^{-4}$ for the constant learning rate $\eta$. The model parameters $\boldsymbol{\theta}$ are randomly initialized with uniform distribution between -1 and 1. % %------------------------------------------------------------------------------- % Fig. 2 \begin{figure}[t] \centering \centerline{\resizebox{1\columnwidth}{!}{\includegraphics{Figs/Dec_vs_eps_RateEnc_ConvFTS_Dec_T_16}}} \caption{{\small{Test accuracy for ML training under adversarial and random changes versus $\epsilon$ with rate encoding for both rate and first-to-spike decoding rules $\left( T = K = 16 \right)$.}}} \label{fig:Dec_vs_eps_RateEnc_ConvFTS_Dec_T_16} \end{figure} %------------------------------------------------------------------------------- % % %------------------------------------------------------------------------------- % Fig. 3 \begin{figure}[t] \centering \centerline{\resizebox{1\columnwidth}{!}{\includegraphics{Figs/Dec_vs_eps_RTEnc_FTS_Dec_T_16}}} \caption{{\small{Test accuracy for ML training under adversarial attacks versus $\epsilon$ with both rate and time encoding rules for first-to-spike decoding $\left(T = K = 16 \right)$.}}} \label{fig:Dec_vs_eps_RTEnc_FTS_Dec_T_16} \end{figure} %------------------------------------------------------------------------------- % We first evaluate the sensitivity of different encoding and decoding schemes to adversarial examples obtained as explained in Sec.~\ref{sec:Adversarial_Examples}. For reference, we consider also perturbations obtained by randomly and uniformly adding, removing and flipping spikes. Fig.~\ref{fig:Dec_vs_eps_RateEnc_ConvFTS_Dec_T_16} illustrates the test accuracy under adversarial and random perturbations when performing standard ML training. The accuracy is plotted versus the adversary's power $\epsilon$ assuming rate encoding and both rate and first-to-spike decoding rules. The results highlight the notable difference in performance degradation caused by random perturbations and adversarial attacks. In particular, adversarial changes can cause a significant drop in classification accuracy even with small values of $\epsilon$, particularly when the most powerful flip attacks are used. First-to-spike decoding is seen to be more resistant to add and flip attacks, while it is more vulnerable than rate decoding to remove spike attacks. The resilience of first-to-spike decoding can be interpreted as a consequence of the fact that the log-likelihood \eqref{FTS_dec_likelihood}, unlike \eqref{Rate_dec_likelihood} for rate decoding, associates multiple outputs to the correct class, namely all of those with the correct neuron spiking first. Nevertheless, removing properly selected spikes can be more deleterious to first-to-spike decoding as it may prevent spiking by the correct neuron. The comparison between rate and time encoding in terms of sensitivity to adversarial examples is considered in Fig.~\ref{fig:Dec_vs_eps_RTEnc_FTS_Dec_T_16} under the assumption of first-to-spike decoding. Time encoding is seen to be significantly less resilient than rate encoding. This is due to the fact that time encoding, in the form considered here of intensity-to-latency encoding, which associated a single spike per input neuron \cite{stromatias2017event}, can be easily made ineffective by removing selected spikes. We then evaluate the impact of robust adversarial training as compared to standard ML. To this end, in Fig.~\ref{fig:Dec_vs_eps_Adv_Tr_eps_5_10_RateEnc}, we plot the test accuracy for the case of flip and remove attacks for both ML and adversarial training when $T = K = 8$. Here we also focus solely on the two classes $\{5, 7\}$. We recall that the adversarial training scheme is parametrized by the time support $T_A$ of the attacks considered during training, here $T_A = 8$, and by its power $\epsilon_A$, here $\epsilon_A = 5/2048$ and $\epsilon_A = 10/2048$. It is observed that robust training can significantly improve the robustness of the SNN classifier, even when $\epsilon_A$ is not equal to the value $\epsilon$ used by the attacker during the test phase. Furthermore, increasing $\epsilon_A$ enhances the robustness of the trained SNN at the cost of a higher computational complexity. For instance, for an attacker in the test phase with $\epsilon = 10/2048$, i.e., with 10 bit flips, conventional ML achieves an accuracy of $45\%$, while adversarial training with $\epsilon_A = 10/2048$ (i.e., 10 bit flips) achieves an accuracy of $87\%$. Finally, the results show that the classifier remains resilient against other type of attacks, despite being trained assuming the flip attack. % %------------------------------------------------------------------------------- % Fig. 4 \begin{figure}[t] \centering \centerline{\resizebox{1\columnwidth}{!}{\includegraphics{Figs/Dec_vs_eps_Adv_Tr_eps_5_10_RateEnc}}} \caption{{\small{Test accuracy under adversarial attacks versus $\epsilon$ with rate encoding and rate decoding with ML and adversarial training $\left( T = K = 8 \right)$.}}} \label{fig:Dec_vs_eps_Adv_Tr_eps_5_10_RateEnc} \end{figure} %------------------------------------------------------------------------------- % Finally, under the same conditions as in Fig.~\ref{fig:Dec_vs_eps_Adv_Tr_t_1_T_TimeEnc_SL}, we study the effect of limiting the power of the adversary assumed during training by considering $T_A = 1$ and $T_A = 8$ with the same $\epsilon_A = 5/2048$. We assume time encoding and rate decoding. It is observed that robust training can still improve the robustness of the SNN classifier, even when $T_A \ll T$ during training. For instance, for an attacker in the test phase with $\epsilon = 5/2048$, i.e., 5 bit flips, conventional ML achieves an accuracy of $34.2\%$, while adversarial training with $\epsilon_A = 5/2048$ and $T_A = 1$ and $8$ achieves accuracy levels of $60.3\%$ and $77.5\%$, respectively. % % ============================================================================== \section{Conclusions}\label{sec:conclusion_sec} % ============================================================================== % In this paper, we have studied for the first time the sensitivity of a probabilistic two-layer SNN under adversarial perturbations. We considered rate and time encoding, as well as rate and first-to-spike decoding. We have proposed mechanisms to build adversarial examples, as well as a robust training method that increases the resilience of the SNN. Additional work is needed in order to generalize the results to multi-layer networks. % ============================================================================== \section{ACKNOWLEDGMENT} %\vspace{-0.4cm} This work was supported by the U.S. NSF under grant ECCS $\#$1710009. O. Simeone has also received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement $\#$725731). % ============================================================================== % %------------------------------------------------------------------------------- % Fig. 5 \begin{figure}[t] \centering \centerline{\resizebox{1\columnwidth}{!}{\includegraphics{Figs/Dec_vs_eps_Adv_Tr_t_1_T_TimeEnc_SL}}} \caption{{\small{Test accuracy under adversarial attacks versus $\epsilon$ with time encoding and rate decoding with ML and adversarial training $\left(T = K = 8 \right)$.}}} \label{fig:Dec_vs_eps_Adv_Tr_t_1_T_TimeEnc_SL} \end{figure} %------------------------------------------------------------------------------- % % ============================================================================== % ============================================================================== % ============================================================================== \bibliographystyle{IEEEtran} \bibliography{ieeetran} % ============================================================================== \end{document} }
\caption{Additional \aastex\symbols}
\caption{The beginning of the document \texttt{functions} referred to by the predator-prey example. The full version is \bluehref{http://khinsen.net/leibniz-examples/examples/functions.html}{available online}.}
\caption{%A standard computational graph is shown in black, while our additions are in {\color{red} red}. %Note that $t$ is the target for some data point $x$, which can be sampled from the training or validation sets. %We want to make predictions $y$ for each data point that are close to the target, with closeness being determined by our loss $\lossSymbol$ and hyperparameters $\hyper$. %The prediction is parameterized by $\param$ which is optimized to minimize the training loss, while the hyperparameters are optimized to minimize validation loss. %This makes the optimal parameters dependent on the hyperparameters and the optimal hyperparameters dependent on the optimal parameters. %We approximate this relationship with a hypernet parameterized by $\responseParam$ and leverage differentiability of the hypernetwork for hyperparameter optimization. \emph{Left:} A typical computational graph for cross-validation, where $\alpha$ are the optimizer parameters, and $\hyper$ are training loss hyperparameters. It is expensive to differentiate through the entire training procedure. % - see \citet{maclaurin2015gradient}. \emph{Right:} The proposed computational graph with our changes in {\color{red} red}, where $\responseParam$ are the hypernetwork parameters. We can cheaply differentiate through the hypernetwork to optimize the validation loss $\lossSymbolOuter$ with respect to hyperparameters $\hyper$. We use $x$, $t$, and $y$ to refer to a data point, its label, and a prediction respectively.}
\caption{ Comparing three approaches to inferring validation loss. \emph{First column:}{\color{red}A Gaussian process}, fit on $\numTrainVs$ hyperparameters and the corresponding validation losses. \emph{Second column:}{\color{green}A hypernetwork, fit on the same $\numTrainVs$ hyperparameters} and the corresponding optimized weights. \emph{Third column:} Our proposed method, {\color{blue}a hypernetwork trained with stochastically sampled hyperparameters}. \emph{Top row:} The distribution of inferred and true losses. The diagonal black line is where predicted loss equals true loss. \emph{Bottom row:} The distribution of differences between inferred and true losses. The Gaussian process often under-predicts the true loss, while the hypernetwork trained on the same data tends to over-predict the true loss. \label{fig:exp5} }
\caption{Successful retrieval examples of CARS196. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Failed retrieval examples of CARS196. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Successful retrieval examples of CUB-200-2011. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Failed retrieval examples of CUB-200-2011. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Successful retrieval examples of Stanford online products. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Failed retrieval examples of Stanford online products. The first image of every row is a query image. Images denoted with \textcolor{red}{x} are false items.}
\caption{Transaction costs in gas without plugins (\textcolor{blue}{\bf blue}), with locking plugin (\textcolor{red}{\bf red}), with transition counter plugin (\textcolor{brown}{\bf brown}), and with both plugins (\textcolor{darkgray}{\bf dark gray}).}
\caption{Transaction costs in gas without plugins (\textcolor{blue}{\bf blue}), with locking plugin (\textcolor{red}{\bf red}), with transition counter plugin (\textcolor{brown}{\bf brown}), and with both plugins (\textcolor{darkgray}{\bf dark gray}).}
\caption{\red{Orientational order parameter: $|\cos\theta|=|\frac{{\bf v}}{{|{\bf v}|}}\cdot \hat{x}|$. Panel (a) shows the ensemble average of $|\cos\theta|$ as a function of the position in the atomistic region. Panel (b) shows the probability distribution of $|\cos\theta|$}.}
\caption{\red{Evolution in time of an instantaneous distribution profile along the trajectory for the ion pairs that are at time, $t=0$, located in the atomistic region (top panel), hybrid region (middle panel) and in the coarse-grained region (bottom panel). Here we consider a GC-AdResS simulation with uncharged coarse-grained model. The results indicate that there is exchange of ion pairs among different regions and it is consistent with the GC-AdResS set up.}}
\caption{\red{Equivalent of Figure \ref{diffnc} for the charged coarse-grained model. As for the uncharged model, also in this case the flux behaves properly.}}
\caption{ (a) Correction to the drag, Eq. (\ref{Eq_M33}), as a function of $\ell_s/\ell_o$ for different $\mbox{Pr}$. Also shown are the three different regimes in Eq.~(\ref{eq:summary}), black solid lines. (b) Comparison between Eq. (\ref{Eq_M33}) and DNS results for Re$=0.05$ by \cite{Yick09} for Pr$=7$ (\textcolor{green}{\LARGE $\circ$}), Pr$=700$ (\textcolor{red}{\LARGE $\circ$}), and by \cite{Magnaudet2018} for Pr$=0.7$ (\textcolor{blue}{$\blacksquare$}) and Pr$=700$ (\textcolor{red}{$\blacksquare$}). Coloured solid lines show Eq.~(\ref{Eq_M33}) for $\epsilon <0.3$, dashed lines for $\epsilon>0.3$. Also shown are power laws in Fr, black solid lines. The vertical dashed line corresponds to Fr=1/Re. }
\caption{Learning curves comparing MVE with learned dynamics ({\color{purple}purple}), MVE without the TD-$k$ trick ({\color{orange}orange}), IB \citep{kalweit} ({\color{blue}blue}), and DDPG ({\color{black}black}) on (a) \textit{cheetah}, (b) \textit{swimmer}, and (b) \textit{walker}. We used $H=10$ for (a,b), but found the same dynamics class inadequate for walker, reducing \textit{walker} experiments to $H=3$ reduces the improvement MVE has to offer over DDPG, but it still exhibits greater robustness to the poor model fit than IB. Note that we use MA-DDPG, not MA-BDDPG in the IB approach. The bootstrap estimation in MA-BDDPG may reduce model use in some cases, so it is possible that MA-BDDPG would have improved performance in \textit{walker}, where the learned dynamics are poor compared to the other environments.}
\caption{{}Conductance evolution during excitation protocol using a sudden change of gate-voltage. The sample is a 43\AA \thick GeBi$_{\text{x}}% $Te$_{\text{y}}$ with "dark" R$_{\square}$=53M$\Omega$ at 4.1K. It was allowed to relax under V$_{\text{g}}$=0V for 40 hours and after recording its "equilibrium" G$_{\text{0}}$ for$\approx$10$^{\text{2}}$s, V$_{\text{g}}$ was swept to -35V within 3.5s producing the G(t) data shown in the figure. The inset illustrates that, after the new V$_{\text{g}}$=-35V is established, the sample conductance decays with a logarithmic time dependence. Dashed line is best fit for a G(t)=G(1)-a\textperiodcentered\log(t) law.}
\caption{Conductance evolution during excitation by a gate protocol using the same parameters as in Fig.15 except that the sample was first put in its PPC-state as per the infrared protocol (Fig.10). The system was then allowed to relax for three hours before changing the gate voltage. The inset illustrates that, after the new V$_{\text{g}}$=-35V is established, the sample conductance decays with a stretched-exponential time dependence. Dashed line is best fit for a G(t)$\propto$exp[-(t/$\tau$)$^{\beta}$] with $\beta$=0.11 and $\tau$=3\textperiodcentered10$^{\text{10}}$s.}
\caption{{\emph{Upper}:} Summary of properties for the Mo-base alloy. For each listed property the gray box refers to the acceptable target properties, the dark gray is the three-sigma uncertainty on the theoretical prediction. The points refer to experimentally measured values with {\color{red}$\times$}~the proposed alloy, {\color{magenta}$\blacksquare$}~MHC, {\color{orange}$\bullet$}~TZC, {\color{green}$\blacklozenge$}~TZM, and {\color{blue}$\blacktriangle$}~ZHM. % {\emph{Lower}:} The compromise between hardness and cost per cycle made in the design of the Mo-base alloy. The white shaded areas show regions that fail to meet hardness and cost targets. The color of shading shows the likelihood of exceeding all of the targets, following the scale on the right. The white circles show the proposed and existing alloys.}
\caption{(Color online) % (a) Secondary electron micrograph image for the Mo alloy. % (b) Hardness as a function of temperature, the black line shows the theoretical prediction and gray the uncertainty. The points refer to experimentally measured values with {\color{red}$\boldsymbol{+}$}~the optimal alloy, {\color{magenta}$\blacksquare$}~MHC, {\color{orange}$\bullet$}~TZC, {\color{green}$\blacklozenge$}~TZM, and {\color{blue}$\blacktriangle$}~ZHM. % }
\caption{% Concrete mixture design instance ($\lambda=1$): B\&B lower bound improvement compared to Gurobi 7.5.2, with a one hour timeout. The B\&B\cref{alg:bb_overview} is labeled BB-$a$-$b$-$c$ where $a$, $b$ and $c$ denote the strong branching lookahead value, the pseudocost initialization approach, and the solver used for lower bounding and solving convex quadratics, respectively. The BB-\textasteriskcentered{} results sort the unexplored nodes in ascending lower bound order. The dashed-dotted line reports best found feasible solution (upper bound). \label{fig:concrete_bnd_evolution_lm_1}}
\caption{% Concrete mixture design instance ($\lambda=1000$): B\&B lower bound improvement compared to Gurobi 7.5.2, with a one hour timeout. The B\&B\cref{alg:bb_overview} is labeled BB-$a$-$b$-$c$ where $a$, $b$ and $c$ denote the strong branching lookahead value, the pseudocost initialization approach, and the solver used for lower bounding and solving convex quadratics, respectively. The BB-\textasteriskcentered{} results sort the unexplored nodes in ascending lower bound order. The dashed-dotted line reports best found feasible solution (upper bound). \label{fig:concrete_bnd_evolution_lm_1000}}
\caption{Additional \aastex\symbols}
\caption{Tomographic reconstruction of human buccal epithelial cell. (a) The captured intensity image of full sensor FOV under circular illumination aperture with $\rho_s$ = 0.65. (b-c) Detailed RI slices at different axial planes of two enlarged regions, see also \textcolor{blue}{Visualization 1}. (d-e) Comparative profile lines of quantitative RI measurement between single stack and OFC method for two selected small regions. The achievable lateral resolution of OFC-ODT technique up to 260 nm. Scale bar denotes 10 $\mu$m and 5 $\mu$m, respectively.}
\caption{\label{fig_xy} %The distance dependence of the shape distortion in four wrongly assumed cosmologies, %assuming a true cosmology of $\Omega_m=0.26$, $w=-1$. %There are four perfect squares with their true shapes plotted in blue dashed lines. %They are measured by an observer located at the origin, and the distances are reprojected in the four wrong cosmologies. %The apparently distorted shapes are plotted in red solid lines. Examples of the rectangular shape distorted by assuming incorrect cosmologies compared to the true fiducial cosmology $\Omega_m=0.26$ and $w=-1$. \red{In 2D comoving coordinates with the observer at the origin, 4 perfect squares are plotted at various distances along one particular line-of-sight direction, in the fiducial model (blue). These squares are then reprojected into am incorrect cosmological model (red), distorting only the radial positions of the corners of each square and resulting in a distorted quadrilateral shapes.} }
\caption{\label{fig_xi} $\hat\xi_{\Delta s}(\mu)$ measured from the SDSS BOSS DR12 galaxies in six redshift bins (three in LOWZ and three in CMASS), assuming the $\Omega_m=0.26$ $\Lambda$CDM cosmology and a more dark energy dominated cosmology with $w_a=-2$. \red{Measurements, without systematic correction, are plotted for each of the six redshift bins and their redshift evolution with respective to the first bin of LOWZ.} In the $w_a=-2$ cosmology, the shapes of $\hat\xi_{\Delta s}(\mu)$ are different from the $\Omega_m=0.26$ cosmology results, and the difference changes with redshift; a large redshift evolution of $\hat\xi_{\Delta s}(\mu)$ is detected in this cosmology, indicating that it is not likely to be the underlying true cosmology of our universe. The measurements in the HR4 mock catalogues (always in the $\Omega_m=0.26$ $\Lambda$CDM cosmology; plotted in green color) match the general shape of curves measured from observational data, indicating that the simulation reproduces the FoG and Kaiser effects. For the $w_a=-2$ cosmology, we also plot the approximate 2PCFs (red dashed lines) inferred using the technique described in Appendix. \ref{sec:approx_2pcf}. The error induced in the approximation procedure is very small. }
\caption{Example excitation diagrams of multi-temperature \HH~populations using our 13 near-IR \HH~lines. For both panels, the total \HH~column density was set to log$_{10}$ $N$(\HH) = 19 and ortho/para = 3. \textit{Top:} 98.5\% of the \HH~population is described by $T$ = 1800 K (red dashed line), and 1.5\% is described by $T$ = 5200 K (blue dashed line) as \cite{Geballe2017} found for part of the BN/KL outflow. The orange line shows a single temperature fit to the 13 points: $T$ = 2263 K and $N$(\HH) = 5.5$\times$10$^{18}$ cm$^{-2}$ with \redchi~= 1.1, indicating a good fit. \textit{Bottom:} 60\% of the \HH~population has $T$ = 1000 K, 30\% has $T$ = 2000 K, and 10\% has $T$ = 3000 K. The orange line shows a single temperature fit to the 13 points: $T$ = 2461 K and $N$(\HH) = 3.5$\times$10$^{18}$ cm$^{-2}$ with \redchi~= 1.3, indicating a good fit. }
\caption{\textit{Top left:} Fitted temperature compared to the distance from Source I in arcseconds (bottom axis) and pc (top axis) assuming a distance of 414 pc. The dashed lines shows the temperature range corresponding to the bottom left plot. Each circle represents a spaxel and is color-coded by the fractional uncertainty in the fitted temperature ($\sigma_{\rm T}$/$T$). The black points represent the median values in each distance bin. \textit{Bottom left:} Similar to the top left plot, but shows a narrower temperature range as indicated by the dashed lines in the top left plot. \textit{Top right:} The reduced chi-square (\redchi) of the fit is shown against the number of \HH~lines used in the fit. \textit{Bottom right:} The fitted temperature is shown against \redchi. }
\caption{High frequency path decomposition of Figure \ref{fig:sim_setup}\textcolor{red}{(May want to combine this and previous figure into subfigures)}}
\caption{\colora{Dataset Characteristics. The evaluation datasets cover a range of distribution types.}}
\caption{\colora{Accuracy of maximum entropy estimates on distributions with varying cardinality. The \msketch is less accurate on discretized datasets, and fails to converge for cardinalities $n < 5$.}}
\caption{Panel detection error $M(\mathcal{T}_P)$ for noise-free \textcolor{blue}{$\mathcal{E}^0$} and underwater \textcolor{DarkGreen}{$\mathcal{E}^*$} environment conditions}
\caption{Position errors \textcolor{red}{$\robotPositionError{S}{M}$} and \textcolor{blue}{$\robotPositionError{S}{F}$}.\\No marker detected for sampling times marked \textcolor{DarkGreen}{green} \label{fig:real_gt_vs_marker_vs_ekf_position}}
\caption{Orientation errors \textcolor{DarkRed}{$\robotOrientationError{S}{M}$} and \textcolor{DarkBlue}{$\robotOrientationError{S}{F}$}.\\No marker detected for sampling times marked \textcolor{DarkGreen}{green} \label{fig:real_gt_vs_marker_vs_ekf_orientation}}
\caption{\label{Figure5} Instantaneous velocity profiles at $\blacksquare$ $t_1$, \textcolor{red}{\textbullet} $t_2$, \textcolor{blue}{$\blacktriangle$} $t_3$ and \textcolor{green}{$\blacktriangledown$} $t_4$ in section 2; top row to bottom row: 50~Hz, 100~Hz and 200~Hz. %and %\textcolor{green}{$\blacklozenge$} $t_5$. Left column axial velocity $U$, right column vertical velocity $W$.}
\caption{\label{Fig-meanvel} Mean velocity profiles at the different sections, from top to bottom $S_1$, $S_2$ and $S_3$, and for different vibration frequencies: $\blacksquare$ no vibration, \textcolor{red}{\textbullet} 50~Hz, \textcolor{blue}{$\blacktriangle$} 100~Hz, \textcolor{green}{$\blacktriangledown$} 200~Hz. Left column axial velocity $\overline{U}$, right column vertical velocity $\overline{W}$. }
\caption{\label{Fig-fluc} Reynolds stresses with different vibration frequencies in section 2, $\blacksquare$ no vibration, \textcolor{red}{\textbullet} 50~Hz, \textcolor{blue}{$\blacktriangle$} 100~Hz, \textcolor{green}{$\blacktriangledown$} 200~Hz.}
\caption{Domain partition. Circles represent degrees of freedom, with \tikzcircle[fill=black]{2pt} denoting points on process $k$, \tikzcircle[fill=gray!50]{2pt} representing \textit{halo} points needed for communication from other processors, and \tikzcircle[fill=white]{2pt} points on other processors.}
\caption{Empirical study on the qualitative comparisons on KITTI 2015 Stereo test set. The figures from left to right correspond to the input left images, estimated disparity maps or error maps by Godard \etal ~\cite{godard2016unsupervised}, block matching, and our method respectively. And the second and fourth rows are the error maps while the estimated disparity maps are plotted above each error maps, the synthesized right views are also presented in the first column. The error map uses the log-color scale described in~\cite{Menze2015CVPR}, depicting correct estimates in \textcolor{blue}{blue} and wrong estimates in \textcolor{red}{red} color tones. Best view in color.}
\caption{\label{fig:model}(Color online) \red{Tetragonal unit cell (lattice parameters $a$, $a$, $c$) and possible superconducting instabilities for a toy model. Crystal fields along the $z$-axis break inversion symmetry. (c)--(f): top view of the four symmetry-allowed superconducting instabilities for both models with uniform, on-site singlet pairing. The color wheel depicts the phase of the superconducting order parameter. The TRS-breaking instability is a linear combination of (e) and (f), which are degenerate. The arrows show the direction of the circulating super-currents within a unit cell in each case.}}
\caption{(Color online) Ginzburg-Landau free energy up to quartic order for our toy model below $T_c$ with $a_{eff}/T_c=-0.9$ and $\beta_2/\beta_1 = 1.5$. a) Two generic TRS\red{-}related degenerate free-energy minima for $\beta_3/\beta_1 = 1.2$ and $\beta_4/\beta_1 = 2.0$. The minima at $(\theta = \pi, \gamma = 0.12\pi)$ and $(\theta = \pi, \gamma = 0.38\pi)$ correspond to left-circulating and right-circulating LSC states respectively with current $I_c$. b) A ring of degenerate free energy minima for $\beta_4 = \beta_2$ and $\beta_3/\beta_1 = 0.9$. }
\caption{(Color online) \red{Structure of the superconducting order parameter at the simplest TRS-breaking instability described in the main text for the Re$_6$(Zr, Hf, Ti) superconductor family. Each sphere represents one of the $12$ symmetrically distinct Re sites within the unit cell. The phase of the order parameter is given by the color wheel.}}
\caption{Performance (PSNR) of the DBPN-SS network and other networks on 4$\times$ and 8$\times$ enlargement. {\color{red}Red} indicates the best performance.}
\caption{Comparison of the DBPN-L and D-DBPN-L on 4$\times$ and 8$\times$ enlargement. {\color{red}Red} indicates the best performance.}
\caption{Quantitative evaluation of state-of-the-art SR algorithms: average PSNR/SSIM for scale factors 2$\times$, 4$\times$ and 8$\times$. {\color{red}Red} indicates the best and {\color{blue}blue} indicates the second best performance. (* indicates that the input is divided into four parts and calculated separately due to computation limitation of Caffe)}
\caption{Schematic phase diagram of the spin-disordered Hubbard chain at large disorder in the the 2D plane labeled by quantum numbers $j$ and $m$. The eta-pairing raising operator $\eta_+$ acts as $\eta_+ |E, S^z_\textrm{total},j,m\rangle = |E+U, S^z_\textrm{total},j,m+1\rangle$ moving eigenstates horizontally in this figure. The particle-hole transformation maps $|E, S^z_\textrm{total}, j, m \rangle$ to $|E-2mU, S^z_\textrm{total}, j, -m \rangle$, equivalent to a mirror symmetry about $m=0$. In the $S^z_\textrm{total} = 0$ sector, the top left corner of the triangle is the vacuum state. \textbf{Region I} (\textcolor{blue}{blue}) are reference states which are destroyed by $\eta_-$ (left edge) or $\eta_+$ (right edge) and contain a mixture of area-law and log-law states. \textbf{Region II} (\textcolor{pink}{pink}) is the region where all eigenstates' entanglement entropies have logarithmic correction due to repeated application of $\eta_+$ (left edge) or $\eta_-$ (right edge). }
\caption{ \textbf{Top:} The SCAEE histograms at reference states corresponding to (from left to right) $j=0,1,2,3$ with $L=8$, subsystem size $l=2$, $S_z=0$ and $W=14$. \textbf{Bottom:} Mean cut-averaged entanglement entropy (CAEE) vs. ln$L_A$ for different reference states [$j=0$ (\textcolor{blue}{blue curve}), $j=1$ (\textcolor{YellowOrange}{orange curve}), $j=2$ (\textcolor{OliveGreen}{green curve}), $j=3$ (\textcolor{red}{red curve})] for the eigenstates with SCAEE value on the left (right) side of the dashed line corresponding to the left (right) figure. }
\caption{ The entanglement entropy difference vs. ln$(L_A)$ for $L=8$, $S_z=0$, and $W=14$ at different quantum number sectors[$j=4$ (\textcolor{black}{black curve}), $j=3$ (\textcolor{YellowOrange}{yellow curve}), $j=2$ (\textcolor{blue}{blue curve}), $j=1$ (\textcolor{red}{red curve})]. The entanglement entropy for each quantum number sector is averaged over the entropy of all the eigenstates in the sector obtained from exact diagonalization. }
\caption{Ensemble averaged von Neumann entropy $S$ for $W=14$ with quarter filled singlon and half filled doublon initial product states. \textbf{First column:} $L=8$, exact diagonalization and periodic boundary conditions. Results are averaged over 400 (top) and 150 (bottom) samples respectively. \textbf{Second column:} $L=12$, TEBD, and open boundary conditions. Results were averaged over 210 samples. For the quarter filling case (\textcolor{blue}{blue curves}), the entropy grows logarithmically with respect to time. For the half filling case (\textcolor{red}{red curves}), the entropy grows as a power law with $t$, with the power law exponent equal to 0.245 for $L=8$, and $0.29$ for $L=12$. \textbf{Third \& Fourth columns:} Ensemble averaged charge imbalance $I$ using the same samples and parameters as the first and the second column respectively. }
\caption{ Entanglement entropy from maximum polarized sector vs. subsystem size $L_A$ for eta-pairing state with $\eta_+^N$ [$N=0$ (\textcolor{blue}{blue curve}), $N=1$ (\textcolor{OliveGreen}{green curve}),$N=2$ (\textcolor{YellowOrange}{orange curve}), $N=3$ (\textcolor{black}{black curve})]. The numerical data is indicated by $*$ and is consistent with the lines generated by Eq.~\eqref{eq:S_max_polarization}. The sample is selected from reference state with average singlon number equal to 0.036, which is far away from the reference state electron number $K=2$. System size $L=8$, polarization $S_z=0$ and disorder strength $W=14$. }
\caption{ Entanglement entropy vs. subsystem size $L_A$ for eta-pairing with $\eta_+^N$ [$N=0$ (\textcolor{blue}{blue curve}), $N=1$ (\textcolor{YellowOrange}{orange curve}), $N=2$ (\textcolor{OliveGreen}{green curve}), $N=3$ (\textcolor{red}{red curve})]. The numerical data is indicated by $*$ and it is consistent with the lines generated by Eq.~\eqref{eq:singlon_EE}. The sample is selected from reference state with average singlon number almost equal to the reference state electron number $K=2$. System size $L=8$, polarization $S_z=0$ and disorder strength $W=14$. }
\caption{ Entanglement entropy difference between different quantum number sectors and their reference states [$j=4$ (\textcolor{blue}{blue curve}), $j=3$ (\textcolor{OliveGreen}{green curve}), $j=2$ (\textcolor{red}{red curve})] vs. ln$[\nu(1-\nu)]$. $\nu$ is defined as $N/(L-K)$, where $N$ is the number of times of applying the $\eta_+$ operators on the reference state and $K$ is the number of electrons in the reference state. The entanglement entropy is linear with respect to ln$[\nu(1-\nu)]$, which agrees with Eq.~\eqref{eq:many_ref}. }
\caption{Mean cut-averaged entanglement entropy (CAEE) vs. $L_A$ for different reference states [$j=0$ (\textcolor{blue}{blue curve}), $j=1$ (\textcolor{YellowOrange}{orange curve}), $j=2$ (\textcolor{OliveGreen}{green curve}), $j=3$ (\textcolor{red}{red curve})]. $L_A$ is the subsystem size. System size $L=8$, polarization $S_z=0$ and disorder strength $W=14$. \textbf{Left:} The CAEE in each reference state sector is averaged over the eigenstates with SCAEE value on the left hand side of the dashed line in Fig.~\ref{fig:SCAEE_edge_sectors}. The mean CAEE in this case indicates a area law. \textbf{Right:} The CAEE in each reference state sector is averaged over the eigenstates with SCAEE value on the right hand side of the dashed line in Fig.~\ref{fig:SCAEE_edge_sectors}. The mean CAEE in this case indicates an sub-volume law. }
\caption{\label{fig:sim} Under the control law \eqref{eq:system}, six agents asymptotically point toward a common point (marked by `\textcolor{red}{o}').}
\caption{Summary of properties for the Ni superalloy. For each listed property the gray box refers to the acceptable target properties, the dark gray is the three-sigma uncertainty on the theoretical prediction. The points refer to experimentally measured values with {\color{red}$\times$}~V210A where measured, {\color{magenta}$\blacksquare$}~refers to Udimet720, {\color{orange}$\bullet$}~LSHR, {\color{green}$\blacklozenge$}~Rene104, and {\color{blue}$\blacktriangle$}~RR1000. % }
\caption{(Color online) % (a) Secondary electron micrograph image. % (b) V210A yield stress as a function of temperature with black the theoretical prediction for the proposed alloy, along with the uncertainty in gray. The points {\color{red}$\boldsymbol{+}$}~show experimental results for the optimal alloy and {\color{blue}$\blacktriangle$}~RR1000. % (c) Oxidation resistance of V210A and RR1000 with temperature. The theoretical predictions are shown in black with uncertainty in gray. % }
\caption{CLEVR-Humans examples showing the model performs novel reasoning skills that do not appear in CLEVR, including: \textcolor{purple}{\textbf{obstructions}}, \textcolor{blue}{\textbf{object uniqueness}}, \textcolor{green}{\textbf{relative distances}}, \textcolor{red}{\textbf{superlatives}} and \textcolor{yellow}{\textbf{new concepts}}. }
\caption{The first five rows show examples of the final attention map produced by the model for CLEVR-Human questions, demonstrating the ability of the model to perform novel reasoning skills and cope with new concepts that have not been introduced in CLEVR. These include in particular: \textcolor{purple}{\textbf{obstructions}}, \textcolor{blue}{\textbf{object uniqueness}}, \textcolor{green}{\textbf{relative distances}}, \textcolor{red}{\textbf{superlatives}} and \textcolor{yellow}{\textbf{new terms}}. The final row shows examples from CLEVR with object occlusions and summation.}
\caption{Recommendation accuracy results by means of Precision / Recall plots for hashtag recommendations. \pred{} and \rec{} outperform the current state-of-the-art in both evaluation settings and in both datasets (\textit{Result 3}).}
\caption{ (a) Schematic drawing of the devised FBG based strain measurement system in the high-resolution mode. \red{Light from t}he broad band light source is filtered with \red{an} optical filter. The filtered light illuminates the FBG whose reflection is detected with an avalanche photodiode. (b) \red{A spectrum illustrating output of the} optical filter method for the FBG based strain measurement. The change of Bragg wavelength of the FBG $\Delta \lambda_{\rm{B}}$ is converted to $\Delta I$ using the optical filter. The change of the intensity of the reflection originates in the shift of \red{wavelength,} $\Delta \lambda_{\rm{B}}$.\label{sche}}
\caption{ \blue{ (a) A photo of a single crystalline sample of volborthite on top of which an optical fiber with FBG is glued using acrylic type glue. The axis of the optical fiber, external magnetic field, and the crystallographic $b$ axis of volborthite orient in the same direction as shown by the arrow \red{in the picture}. (b) Transmittance spectrum of the optical filter used in the optical filter method. (c) Spectra of reflected light from FBG under various conditions as is indicated in the legend. All spectra are measured with an infrared spectrometer except for spectrum (2) which is measured using an optical spectrum analyzer (OSA) shown for comparison. } \label{spectra} }
\caption{ \blue{ (a) Time evolution of the intensity change of the reflected light from the FBG. (b) The waveform of a pulsed high magnetic field with a 35 ms time duration. (c) Longitudinal magnetostriction of volborthite deduced using the data in (a). Inset shows the magnification of the magnetostriction at the 1/3 magnetization plateau region. (d) The magnetization curve of volborthite. \cite{yoshida} } \label{volbo}}
\caption{Left: components $\eta_R$ ($\blacksquare$) and $\eta_I$ ($\square$) of the visco-elastic modulus $\tilde \eta$ of the solutions as a function of polymer concentration, at 30 Hz (black), 220 Hz (red), and 248 Hz (blue). Right: interfacial friction coefficient $\tilde \lambda = \lambda_R$ at the solution/solid boundary as a function of the HPAM concentration, at 30Hz ({\Large${\bullet}$}), 220Hz (\textcolor{red}{{\Large$\bullet$}}), and 248 Hz (\textcolor{blue}{{\Large$\bullet$}}). The dashed lines are a guide for the eye. }
\caption{Top graph: real value of the viscosity $\eta_R$ as a function of the HPAM concentration ($\circ$) 30 Hz, (\textcolor{red}{$\circ$}) 220 Hz, (\textcolor{blue}{$\circ$}) 248 Hz. The dashed-lines are best power-law fits in $c^{3/2}$, $\eta_R=\eta_s+Ac^{3/2}$. The prefactor $A$ depends on frequency. Bottom graph: real value of the shear modulus $G_R$ at different frequencies ($\square$) 30 Hz, (\textcolor{red}{$\square$}) 220 Hz, (\textcolor{blue}{$\square$}) 248 Hz. The dashed-lines are best power-law fits in $c^{3/2}$, $G_R=Ac^{3/2}$.}
\caption{Friction coefficient at the HPAM solutions/solid interface as a function of the concentration at different frequencies: 30Hz ({\Large${\bullet}$}), 220 Hz, (\textcolor{red}{\Large${\bullet}$}) 248 Hz, (\textcolor{blue}{\Large${\bullet}$}) 248 Hz. }
\caption{Thickness of the interfacial layer of water $e$ obtained from the two-fluid model, as a function of the concentration of the solution C in g/l: ($\blacksquare$) 30 Hz, (\textcolor{red}{$\blacksquare$}) 220 Hz, (\textcolor{blue}{$\blacksquare$}) 248 Hz. The dashed black lines plots the power-laws $y=(\xi+2l_D) C^{-1/2}$ with $\xi=23 $ nm and $l_D = 5.8$ nm. }
\caption{{\color{blue}Summary of the comparison with previous work.}}
\caption{{\small Comparison between state-of-the-art transfer learners and BEETLE. The best transfer learner is shaded \colorbox{lightergray}{gray}.}}
\caption{{\color{myc1}{Total power consumption with $N=10$.}} }
\caption{{\color{myc1}{Total power consumption with respect to $L$ with $N=5$. }}}
\caption{Visualizations of a subset of nodes from \cocitation{} graph with selected conferences: \protect\tikz\protect\draw[cycle2,fill=cycle2] (0,0) circle (.7ex);~VLDB, \protect\tikz\protect\draw[cycle5,fill=cycle5] (0,0) circle (.7ex);~ICDE, \protect\tikz\protect\draw[cycle3,fill=cycle3] (0,0) circle (.7ex);~KDD, \protect\tikz\protect\draw[cycle1,fill=cycle1] (0,0) circle (.7ex);~WWW, and \protect\tikz\protect\draw[cycle4,fill=cycle4] (0,0) circle (.7ex);~NIPS\@. Note that the number of nodes per class is the same for all conferences.}
\caption{FSM truth table for general synchronizer with save depth $D$. FSMs are initialized to $S_D$. \bluecomment{TODO: fix this}}
\caption{ Overview of each cluster and the features they present. Columns represent the features (I: connection to Users' Social Networks, II: Instant Messaging Service, III: In-app Sharing, IV: Friend-Inviting Program). Row is the cluster. \checktikz means the cluster has that particular feature. }
\caption{Analysis of electron diffraction patterns of graphene from experiment (Fig.~\ref{fig:diffraction}a) and different simulation methods (including IAM parameterizations by Kirkland~\cite{Kirkland98}, Lobato~\cite{Lobato14ACA}, Weickenmeier~\cite{Weickenmeier91ACA}, Peng~\cite{Peng96ACA}, Rez~\cite{Rez94ACA}, and directly from GPAW). {\color{red}$\bigcircle$} = 1}
\caption{Analysis of electron diffraction patterns of hBN from experiment (Fig.~\ref{fig:diffraction}b) and different simulation methods. {\color{red}$\bigcircle$} = 1}
\caption{\label{fig:glproportions2}The number of states with gridlock probabilities over the total number of states against the global density $\rho^{(5)}_{\text{global}}$ for different values of $\hat{L}_{153}/ \hat{L}_{14}$. To calculate the ratio the $\left(n_{\text{l}}^{(j_1)},~n_{\text{l}}^{(j_2)}\right)$ - landscapes were discretized in steps of 0.01 and the number of states with gridlock possibility were counted and then divided by the total number of states.} \end{figure} In Figure~\ref{fig:glproportions2} the ratio of the number of states with gridlock possibility over the total number of states is shown against the global density for various pathlength ratios for $L_{1}=100$ and $L_{2}=500$. To obtain this plot, the $\left(n_{\text{l}}^{(j_1)},~n_{\text{l}}^{(j_2)}\right)$ - landscapes were discretized in steps of 0.01 and the number of states with gridlock possibility were counted. Since in our case $\hat{L}_{153}$ is always smaller than $\hat{L}_{14}=\hat{L}_{23}$, the lowest density with a gridlock probabiliy is always the density with $M=\hat{L}_{153}$ (cf. Eq.~(\ref{eq:gridlock153-3})). One can see that for lower values of $L_5$ a large region of the phase space is comprised of states with gridlock probabilities even for low densities (e.g. over 60\% of states can become gridlocked at densities of $\rho^{(5)}_{\text{global}}\approx 0.4$ for $\hat{L}_{153}/ \hat{L}_{14}=0.4$). For longer $E_5$ less states can become gridlocked. Please note that this figure depends on how fine the discretization of space is chosen and also on how phase space is desribed. It might look different if the phase space is chosen to be three-dimensional and described by $(N_{14},~N_{23},~N_{153})$ instead of $\left(n_{\text{l}}^{(j_1)},~n_{\text{l}}^{(j_2)}\right)$. When a gridlock state will occur in the system may depend on the individual realization of the stochastic process. A closer look at when gridlocked states occur is given in Section~\ref{sec:hatched}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Monte-Carlo observables} \label{sec:mc-observables} We analysed the system by employing Monte Carlo (MC) simulations. Details of the measurement processes can be found in~\ref{sec:app-measurement-process}. In a first step, the system and user optima were found by analysing the values of the two observables \begin{eqnarray} \Delta T&=&|T_{14}-T_{23}|+|T_{14}-T_{153}|+|T_{23}-T_{153}|, \label{eq:deltat} \\ T_{\text{max}}&=&\text{max}[T_i, i\in \{14,23,153\} ]. \label{eq:tmax} \end{eqnarray} $T_i$ denotes the travel time on route $i$, i.e. the time a particle needs to go from $j_1$ to $j_4$ via this route. The user optima and system optima are given by the combinations of $N_{14}$, $N_{23}$, $N_{153}$ (or $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$) which minimize $\Delta T$ and $T_{\text{max}}$, respectively. A true user optimum is characterized by $\Delta T(n_{\text{l}}^{(j_1)}, n_{\text{l}}^{(j_2)})=0$ since then all routes have the same travel times. If there are any unused routes and the travel times of these unused routes are higher than that of the used ones, $\Delta T$ is reduced to only the absolute values of the travel time difference of the used routes (see~\ref{sec:app-special} for an example of this case). In our simulations, a strategy with a $\Delta T$ value close to zero is regarded as a user optimum since the exact zero case is seldom found when measuring travel times. The system optimum is given by the strategy which minimizes $T_{\text{max}}$. By comparing the positions and travel time values of the user and system optima of the 4link and 5link systems, the phase diagram can be constructed according to the scheme described in Section~\ref{sec:possilbe-states}. If a true user optimum is found, all three routes have the same travel time. In this case this travel time can be compared to the maximum travel time of the system optimum. In cases where we could not find exact user optima, we compared the maximum travel time of the 'closest candidate' for the user optimum (state with lowest $\Delta T$) to the maximum travel time of the system optimum. More details on the cases where no real user optima could be found are given in Section~\ref{sec:hatched}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Results} \label{sec:results} \subsection{Travel times in the 4link network} \label{sec:results_4link} Since the 4link system, our reference system, is symmetric one expects the user optimum and the system optimum to be given by half the particles taking route 14 and the other half taking route 23 for all possible global densities $\rho_{\text{global}}^{(4)}$. In Fig.~\ref{fig:t4link} we can see that this is indeed the case. \begin{figure}[h!] \centering \includegraphics{./t4link.pdf} \caption{\label{fig:t4link}Results of MC simulations of the 4link system for 50\% of the particles choosing route 14 and 50\% choosing route 23 for the whole density regime $0\leq \rho_{\text{global}}^{(4)} \leq 1$ and $L_1=L_3=100$ and $L_2=L_4=500$. Part \textbf{(a)} shows, that throughout all densities, the travel times on routes 14 and 23 are almost equal. Part \textbf{(b)} shows the average travel times on these routes (\textcolor{orange-4link}{orange $\times$'s}) and for comparison the travel time of a single TASEP with $M/2$ particles according to Equation~(\ref{eq:tperbc}) (\textcolor{blue-4link}{blue line}). One can see, that jamming at $j_4$ plays an important role for densities $\rho_{\text{global}}^{(4)}\gtrsim 0.2$. As can be seen in part \textbf{(c)}, the relative standard deviations of the travel times on both routes are below $5\%$ for all densities (values for route 14/23 in \textcolor{red-4link}{red}/\textcolor{green-4link}{green}.} \end{figure} In Figure~\ref{fig:t4link}~\textbf{(a)} we see the value of $\Delta T=|T_{14}-T_{23}|$ (Equation~(\ref{eq:deltat}) reduces to this form in the 4link system) against the global density. The value is close to zero for all global densities which means that the travel times are (almost) equal on both routes and this symmetric distribution of the particles is indeed the user optimum. Since the network is symmetric, this is also the system optimum, as any unequal distribution of the particles would lead to a higher travel time on the route with more particles. In Figure~\ref{fig:t4link}\textbf{(b)} the average of the travel times measured on routes 14 and 23 ($T_{av}=(T_{14}+T_{23})/2$) is shown. For comparison also the travel time of a single TASEP used by $M/2$ particles (obeying Equation~(\ref{eq:tperbc})) is indicated by the blue line. One can see that for densities $\rho_{\text{global}}^{(4)}\gtrsim 0.2$ the jamming at $j_4$ plays an important role since the travel times are higher than those of the single TASEP from this density upwards. In Figure~\ref{fig:t4link}\textbf{(c)} we can see the relative standard deviation of the travel time measurements of both routes \begin{equation} \sigma(T)/T=\frac{1}{\bar{T}}\left(\frac{1}{n}\sum_{i=1}^n (T_i-\bar{T})^2\right)^{1/2}\label{eq:sd} \end{equation} with $n$ being the number of measurements and \begin{equation} \bar{T}=T=\frac{1}{n}\sum_{i=1}^nT_i \end{equation} being the mean of all measured values. One can see that it stays below $5\%$ for all densities. This means we measure stable travel time values for all densities which is also supported by the measured density profiles (Figure~\ref{fig:rho4link}). They show a sharp domain wall between low and high density regions at a fixed position which is the same on both routes. \begin{figure}[h!] \centering \includegraphics{./4link-densities-L0-1-L13-100-L24-500vergleiche.pdf} \caption{\label{fig:rho4link}Density profiles of the two routes in the 4link system with $L_1=L_3=100$ and $L_2=L_4=500$ for the four different global densities $\rho_{\text{global}}^{(4)} \in \{0.1,0.2,0.5,0.75\}$ printed in $\{$\textcolor{purple-density}{purple}, \textcolor{orange-density}{orange}, \textcolor{brown-density}{brown}, \textcolor{blue-density}{blue}$\}$. The local densities $\rho_i^{14/23}$ on the two routes are shown against the position $i$. The denisity on $E_0$ is given by a $+$, the density of junction sites on the roads by $\times$'s. One can see that in all cases the density profiles are almost equal on both routes. Domains walls form at the same fixed positions on both routes.} \end{figure} This is a very important difference to the model with higher stochasticity due to turning probabilities~\cite{PhysRevE.94.062312}. In the more stochastic case, fluctuating domain walls dominate the density region $0.29\lesssim\rho^{(4)}_{\text{global}}\lesssim0.75$ resulting in (short-term) unstable travel time values. A comparison of the different behaviours of the two models is shown in Figure~\ref{fig:t4linkcomparison}. \begin{figure}[h!] \centering \includegraphics{./t4link-comparison-numbers.pdf} \includegraphics{./t4link-comparison-probabilities.pdf} \caption{\label{fig:t4linkcomparison}A comparison of the density profiles of routes 14 and 23 in the 4link systems with $L_1=L_3=100$ and $L_2=L_4=500$ and $M=622$, $\rho_{\text{global}}^{(4)}\approx 0.52$ in the present less stochastic model (part \textbf{(a)}) and in the model with higher stochasticity (with turning probability $\gamma =0.5$) treated in our previous article~\cite{PhysRevE.94.062312} (part \textbf{(b)}). In both parts, the averaged density profiles over the whole measurement time of $10^6$ sweeps is shown in \textcolor{red}{red} and 50 different shortterm instances (averaged over $10^4$ sweeps each) are shown in 50 different shades of gray. The density on $E_0$ is given by a $+$, the density of junction sites on the roads by $\times$'s. One can see that in the present model, the domain wall position is fixed, while it fluctuates through the routes in a coupled manner in the model with higher stochasticity.} \end{figure} In Fig.~\ref{fig:t4linkcomparison} \textbf{(a)} and \textbf{(b)} density profiles of the present model and of that treated in~\cite{PhysRevE.94.062312} are shown. The averaged density profiles measured over the whole measurement time of $10^6$ sweeps are given in red. In different shades of gray, different measurement instances (averaged over $10^4$ sweeps each) are shown. As one can see in part \textbf{(a)} of the figure, in the present model the domain wall is at the same fixed position during the whole measurement process: all the grey curves lie below the red curve. In the model with turning probabilities this is not the case as seen in part \textbf{(b)}. Here, the averaged density profile becomes close to a straight ascending line on both paths. This is due to the fact that the domain walls perform a coupled random walk on the two paths. If the high density region gets longer on one path it gets shorter on the other one respectively. The domain wall positions can be at any point of the paths depending on the measurement time. Therefore, particles entering the same route at different times may encounter a completely different situation and thus have completely different travel times even if the system is in its stationary state and none of the parameters changes. This behaviour of the model with higher stochasticity is treated in detail in~\cite{PhysRevE.94.062312}. Due to the fluctuating domain walls and the resulting unstable travel times in the model with higher stochasticity $uo$ and $so$ could not be determined in a large intermediate global density region. Thus the system's reaction to the addition of $E_5$ could not be classified according to our scheme presented in Section~\ref{sec:possilbe-states} in this density region. In the present case we obtain stable travel time values and thus $uo$ and $so$ throughout the whole density region for the 4link system. Thus we are able to compare the 5link system to the 4link system in the whole density region. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Phase diagram} We used Monte Carlo simulations to obtain the user optima and system optima of the 5link system for different combinations of $L_5$ and $M$. States with potential gridlock formation as explained in Section~\ref{sec:gl} were not considered as candidates for user or system optima. Even if some of these states may, depending on the initial configuration and realization of the stochastic process, have values of $\Delta T$ close to zero in the beginning of the system's time evolution, they all evolve into a gridlock steady state (more details on this in Sec.~\ref{sec:hatched}). By comparing the travel times of the found 5link user and system optima to those of the 4link system's user and system optima for the same $M$ we could derive the phase diagram of the system according to the classification shown in Figure~\ref{fig:possiblestates}. In~\ref{sec:app-measurement-process} we describe our procedure to find the user and system optima in more detail. The phase diagram (Fig.~\ref{fig:phases}) and the following Fig.~\ref{fig:closerlook} show the influence of the control parameters $L_5$ and $M$: The $x$-axis is always given by $\hat{L}_{153}/\hat{L}_{14}$ which is the ratio of the lengths of the new route 153 and the two old routes 14 and 23 (see Eqs.~(\ref{eq:l153-1}) -- (\ref{eq:l153-2})). There are always two $y$-axes which decode the number of particles $M$ via the global densities in the 4link/5link systems $\rho_{\mathrm{global}}^{(4)/(5)}$ (see Eqs.~(\ref{eq:rho4}) -- (\ref{eq:rho4-5})). %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsubsection{Phase diagram's structure} The phase diagram shown in Figure~\ref{fig:phases}~\textbf{(a)} can be divided into three parts. \begin{figure}[h!] \centering \includegraphics[width=0.49\columnwidth]{./phases.pdf} \includegraphics[width=0.49\columnwidth]{./deltaT.pdf} \caption{\label{fig:phases} The shown results were obtained for $L_1=L_3=100$, $L_2=L_4=500$ and varying lengths of $E_5$ (resulting in the x-axis $\hat{L}_{153}/\hat{L}_{14}$) and $M$ (resulting in the two y-axes $\rho_{\mathrm{global}}^{(4)/(5)}$). \textbf{(a)} The phase diagram according to the classification scheme given in Figure~\ref{fig:possiblestates}. The $\times$'s show where simulations were performed. The phase boundaries are drawn according to those simulations and are thus, due to this limited number of simulations, only a rough estimate of the phase boundaries. In phases A, B and C real user optima could by found while in the area marked by a hatching (phases B$^{\prime}$, C$^{\prime}$, D$^{\prime}$, E$^{\prime}$) no real user optima could be found. In phase F the 4link system is full. \textbf{(b)} The value of $\Delta T$ for selected measurement points. One can see that it is below 100 (indicated by coloured $\bigcirc$) in the unhatched area and above 100 (indicated by coloured $\bigtriangleup$) in the hatched area.} \end{figure} The trivial first part is called phase F. This phase is present for really high densities in the 5link system. In this phase, there are more particles in the 5link system than sites in the 4link system. The 4link system is thus completely blocked and the two systems cannot be compared. One could argue that $E_5$ improves the system at these high densities, but only in the sense that the 4link could not even take up that many particles. The second part, consisting of phases A$_{1/2}$, B and C, is the part in which real user optima of both the 4link and the 5link system exist in the sense that states in which $\Delta T \approx 0$ exist, i.e. states with almost equal travel times on all three routes. The third part is given by the region above the light dotted line which is marked by a hatching and consists of the "primed" phases B$^{\prime}$, C$^{\prime}$, D$^{\prime}$ and E$^{\prime}$. In all of these phases no real user optima exist in the sense that no states with $\Delta T<100$ could be identified. This means that in neither of these phases a distribution of the particles onto the routes exists which leads to (almost) equal travel times on all routes. This is a consequence of gridlocked states: Combinations of $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$ which would potentially be a user optimum are not available since they would lead to a gridlock. The values of $\Delta T$ in some points of the phase diagram are shown in Figure~\ref{fig:phases}~\textbf{(b)}. As can be seen, in phases A, B and C the value is zero or of the magnitude of 10, while it is higher than 100 in the hatched phase, reaching up to values of the order of $10^4$ at really high densities. Phases A$_{1/2}$ are "$E_5$ optimal" phases. In these phases, the travel times in the $uo$ of the 5link system are lower than that of the $uo$ in the 4link system and $uo^{(5)}=so^{(5)}$. In phase A$_1$, the region below the dotted line at low path length ratios and low global densities, the $uo$ of the 5link is given by all particles choosing route 153. The new route is the only used route of the system. In phase A$_2$, above the dotted line, all three routes are used in the 5link user optima. Phase B is the "Braess 1" phase. Here the $so$ of the 5link is equal to that of the 4link, meaning half of the particles choose route 14 and the other half route 23. The user optimum of the 5link is given by another distribution of particles onto the three routes, such that all three routes have the same \textit{higher} travel time. The "Braess 1" phase actually dominates the largest part of the phase diagram. There is also a very small part of the phase diagram at densities around $\rho_{\mathrm{global}}^{(4)}\approx 0.8$ and $\hat{L}_{153}/\hat{L}_{14}\gtrsim 0.7$, in which the "Braess 2" phase (phase C) is found, meaning that with traffic regulations the system could be improved due to $E_5$ ($T_{\mathrm{max}}(so^{(5)})<T_{\mathrm{max}}(so^{(4)})$), but without regulations - in the $uo$ - the travel times are increased due to the addition of $E_5$. In the hatched area, in the 5link system no real user optima with equal travel times on all routes exist. This is because combinations of $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$ which could potentially lead to equal travel times on all routes lead to gridlocks in the system according to Section~\ref{sec:gl}. Only states without the possibility of a gridlock were taken into account for the construction of the phase diagram. In most cases, the state with the lowest value of $\Delta T$ and without gridlock possibility is a state in which the travel time on route 153 is lower than on the other two. When more particles choose route 153, $T_{153}$ increases. But before a state with equal travel times on all routes is reached, route 153 and subsequently all three routes come to a complete gridlock. More details on the hatched area and an illustrative example can be found in Section~\ref{sec:hatched}. For the construction of our phase diagram we used the state with the lowest value of $\Delta T$ and without gridlock possibility we could find and compared its maximum travel time to that of the system optimum and the user optimum of the 4link system. It is important to keep in mind that even though in this manner we could identify different phases inside the hatched region, there are no real user optima in the 5link in all of the primed phases. In a system with real drivers more and more drivers would tend to switch routes with the desire to reduce their own travel time and in the long run the system will evolve into a complete gridlock. Here, judging from the maximum travel times in the state with the lowest $\Delta T$, we could identify two phases in which the system could not be improved even if the traffic was externally regulated. These are the "$E_5$ not used - like" phase (phase A$^{\prime}$) and the a "Braess 1 - like" phase (phase B$^{\prime}$). In both of those phases the $so$ of the 5link equals that of the 4link. While in the former this is also the state with the lowest value of $\Delta T$ (and real drivers would thus tend to switch to route 153), in the second one another state with a higher maximum travel time gives the lowest value of $\Delta T$. In the "Braess 2 - like" phase (phase C$^{\prime}$) the system could be improved by the addition of $E_5$ if the traffic is regulated but is not for selfish drivers. In the "$E_5$ improves - like" phase (phase E$^{\prime}$), the maximum travel time in the state with the lowest $\Delta T$ in the 5link is lower than that of the 4link but is not as low as in the 5link's system optimum. Summarizing we can conclude that in a system of selfish drivers without traffic regulations the addition of $E_5$ only reduces travel times at low densities $\rho_{\mathrm{global}}^{(5)}\lesssim 0.2$ (phases A$_{1/2}$). Here the system with the new road will also be in its optimal state ($uo^{(5)}=so^{(5)}$). At really high densities, the system shows "$E_5$ improves - like" behaviour (phase E$^{\prime}$), meaning that in the 5link system the state in which all three travel times are closest to each other has a lower maximum travel time than the user optimum of the 4link system. Still the 5link is not in its optimum in this state and, more importantly, this is not a real user optimum. Instead the system is prone to gridlocks when used by real drivers since more and more of them would tend to choose route 153. In most parts of the system, the addition of $E_5$ leads to higher travel times in the system used by selfish drivers (phases B, B$^{\prime}$, C$^{\prime}$) or the new road will be ignored (phase D$^{\prime}$). In most of these phases (phases B, B$^{\prime}$, D$^{\prime}$) even if external traffic guidance was applied, $E_5$ would not reduce travel times. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsubsection{The region without proper user optima} \label{sec:hatched} Here we take a closer look at the phases in the area of the phase diagram which is marked by a hatching (Figure~\ref{fig:phases}) by analysing an exemplary point of this region: we look at the point with the parameters $\hat{L}_{153}/\hat{L}_{14}=0.5$ and $\rho_{\mathrm{global}}^{(5)}=0.49$. According to the phase diagram shown in Fig.~\ref{fig:phases} this is a point of the B$^{\prime}$ phase ("Braess1 - like" phase). In Fig.~\ref{fig:hatchedexample} we show the $n_{\text{l}}^{(j_1/j_2)}$ - landscapes and the corresponding $\Delta T$ and $T_{\mathrm{max}}$ values. In Fig.~\ref{fig:hatchedexample} certain points are marked and the corresponding travel time values of the three paths and the $\Delta T$ and $T_{\mathrm{max}}$ values for these points are shown in Table~\ref{tab:hatched-exp}. \begin{figure}[h] \centering \includegraphics[width=0.49\columnwidth]{./meassurements-and-gridlock-theoretical-L0-1-L13-100-L24-500-L5-97-M-638-uo.pdf} \includegraphics[width=0.49\columnwidth]{./meassurements-and-gridlock-theoretical-L0-1-L13-100-L24-500-L5-97-M-638-so.pdf} \caption{\label{fig:hatchedexample}The $n_{\text{l}}^{(j_1/j_2)}$ - landscapes of \textbf{(a)} the $\Delta T$ and \textbf{(b)} the $T_{\mathrm{max}}$ -values for $L_5=97$ and $M=638$, which means $\hat{L}_{153}/\hat{L}_{14}=0.5$ and $\rho_{\mathrm{global}}^{(5)}=0.49$. States with a gridlock possibility on routes 14, 23 and 153 are marked with \textcolor{blue-gl}{blue $\times$'s}, \textcolor{red-gl}{red~$+$'s} and \textcolor{green-gl}{green $\star$'s} respectively. Simulations were performed for regions where no gridlock is possible and $n_{\text{l}}^{(j_1/j_2)}$ was sweeped in steps of 0.01. The pink point $+1$ is the point with the lowest value of $\Delta T$. Thus this is the point which is used for the phase diagram. The grey points $+2$ and $+3$ have smaller values of $\Delta T$ but route 153 can gridlock. The white point $\star 1$ has the lowest value of $T_{\mathrm{max}}$ and is thus the system optimum.} \end{figure} In part \textbf{(b)} of Fig.~\ref{fig:hatchedexample} the point $\star 1$ represents the system optimum of this parameter set. It is given by half the particles choosing route 14 and the other half route 23. This point has the lowest value of $T_{\mathrm{max}}=1789$. It is not the user optimum as it has a high value of $\Delta T=2230$ since the travel time on route 153 is much lower than on the other routes. More and more particles would tend to switch to route 153. The point $+1$ shown in Fig.~\ref{fig:hatchedexample}~\textbf{(a)} is the point which we used for our construction of the phase diagram. It is the point with the lowest value of $\Delta T=2215$ of all the points without gridlock possibility. From Table~\ref{tab:hatched-exp} we see that at this point route 153 still has a much lower travel time than the routes 14 and 23. In a system with real selfish drivers, more and more drivers would thus switch onto route 153. \begin{table}[h] \caption{\label{tab:hatched-exp} The $n_{\text{l}}^{(j_1/j_2)}$, $\Delta T$, $T_{\mathrm{max}}$ and travel time values of the routes for the 4 points which are marked in Figure~\ref{fig:hatchedexample}~\textbf{(a)} and ~\textbf{(b)}. Point $+ 1$ is the point with the lowest value for $\Delta T$ without a gridlock possibility. Thus this is the point we chose for construction of the phase diagram. Point $\star 1$ is the system optimum. The points $\times 1$ and $\times 2$ are states with gridlock possibilities on route 153. The measured travel time values are marked with a star because they were measured before the system gridlocked and are not stationary state values.} \begin{center} \begin{tabular}{| c | l | l | l | l | l | l | l |} \hline Point & $n_{\text{l}}^{(j_1)}$ & $n_{\text{l}}^{(j_2)}$ & $T_{14}$ & $T_{23}$ & $T_{153}$ & $\Delta T$ & $T_{\mathrm{max}}$ \\ \hline $+ 1$ & 0.730 & 0.798 & 2270 & 2047 & 1162 & 2215 & 2270 \\ \hline $\times 2$ & 0.741 & 0.735 & 2113$^{\star}$ & 2015$^{\star}$ & 1539$^{\star}$ & 1148$^{\star}$ & 2113$^{\star}$ \\ \hline $\times 3$ & 0.752 & 0.671 & 2797$^{\star}$ & 2744$^{\star}$ & 2748$^{\star}$ & 106$^{\star}$ & 2797$^{\star}$ \\ \hline\hline $\star 1$ & 0.5 & 1.0 & 1789 & 1789 & 674 & 2230 & 1789 \\ \hline \end{tabular} \end{center} \end{table} From Figure~\ref{fig:hatchedexample}~\textbf{(a)} we know that if more particles choose route 153, a gridlock on that route becomes possible. We marked two more points (point $\times 1$ and $\times 2$) in this figure. From Table~\ref{tab:hatched-exp} we see that the value of $\Delta T$ decreases for those two points. It is important to keep in mind though that the travel time values of points $\times 1$ and $\times 2$ were measured before the system gridlocked. In Figure~\ref{fig:hatchedpoint3} we see how travel times of the individual routes and the value of $\Delta T$ develop during the measurement process at point $\times 3$. \begin{figure}[h] \centering \includegraphics{./tentwicklung-various-rng-seeds-L0-1-L13-100-L24-500.pdf} \caption{\label{fig:hatchedpoint3}The time evolution of \textbf{(a)} - \textbf{(c)} the measured mean values of the travel times of the three routes and \textbf{(d)} the value of $\Delta T$ for the state $\times 3$ shown in Figure~\ref{fig:hatchedexample}. The different coloured curves show six different instances of the system realized with six different seed values for the random number generator. The second y-axes on the right sides of the first three figures show how many times $n_i$ the values of $T_i$ were measured, represented by the dotted grey line. The vertical lines show the points in time when the system gridlocked on route 153.} \end{figure} For this figure six instances of the system with different seed values for the random number generator were generated and the travel times were measured during the evolution of the system. The system was not relaxed before measurements begun (the system was always relaxed for all measurements which were used for the phase diagram though - compare~\ref{sec:app-measurement-process} for details on our general measurement process). The relaxation was skipped here since otherwise the system may have already gridlocked during the relaxation process. One can see that the state $\times 3$ is actually a good candidate for a user optimum since the value of $\Delta T$ seems to be very low since all three routes have similar travel times (also compare Table~\ref{tab:hatched-exp} for the numbers). All six instances of the system gridlock at some point of the time evolution. The earliest gridlock occurs after 130000 sweeps (blue line) and the latest after 1470000 sweeps (grey line). This example shows rather clearly that in the hatched area of the phase diagram, if the system was actually used by intelligent selfish particles, the gridlock possibility is very high. This is because states with gridlock possibilities have actually lower possible values of $\Delta T$. Thus more and more particles would tend to switch routes with the aim of reducing their own travel times but with the risk of causing a gridlock of the whole network. This is why for our phase diagram (Figure~\ref{fig:phases}) and for our further analysis on the influence of $E_5$ on the system (see Sec.~\ref{sec:closerlook} and Fig.~\ref{fig:closerlook} therein) we only considered points without gridlock possibility and measured the travel time values of the steady states. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsubsection{A closer look at the influence of the new edge} \label{sec:closerlook} When talking about the Braess paradox, foremost one is interested in the relation of the travel times of the user optima of a road network before and after adding a new road. In Figures~\ref{fig:closerlook} \textbf{(a)} to \textbf{(c)}, our two systems are studied in a bit more detail by comparing $so^{(5)}$ to $so^{(4)}$, $so^{(5)}$ to $uo^{(5)}$ and finally $uo^{(5)}$ to $uo^{(4)}$. Note that in parts \textbf{(a)} and \textbf{(b)} of the figure, the values used for $T_{\mathrm{max}}(so^{(5)})$ may actually not be the absolutely lowest possible maximum travel times. This is due to the fact that, opposed to the value of $\Delta T$ going to zero in the user optimum, a natural lower bound for $T_{\mathrm{max}}$ does not exist. Therefore one cannot be sure if the actual system optimum was found precisely (for more details, see~\ref{sec:app-sweep}). Furthermore it is not guaranteed that we found all possible user optima and therefore the values (given by the colors) should be interpreted as tendencies of how the new road influences the system and not as exact values (see~\ref{sec:app-special} for details - in some cases actually multiple $uo$s were found). In Figure~\ref{fig:closerlook} \textbf{(a)} additionally to the underlying phase structure, for some specific points, the ratio of the system optimum's maximum travel time in the 5link system and that of the 4link system $T_{\text{max}}(so^{(5)})/T_{\text{max}}(so^{(4)})$ is shown. \begin{figure}[h] \centering \includegraphics[width=0.49\columnwidth]{./max-so-4-to-so-5.pdf} \includegraphics[width=0.49\columnwidth]{./max-uo-5-to-so-5.pdf} \includegraphics[width=0.49\columnwidth]{./max-uo-4-to-uo-5.pdf} \caption{\label{fig:closerlook}Quantifications of the influence of $E_5$ on the network. In all three parts, white $\bigcirc$ indicate values equal to 1, coloured $\bigtriangleup$ indicate values between 1 and 1.5, coloured $\bigtriangledown$ indicate values between 0.5 and 1, green $\diamondsuit$ indicate values below 0.5 and yellow $\Box$ indicate values above 1.5. Part \textbf{(a)} shows how the new road could improve the system if it was always in its system optimum, measured by $T_{\text{max}}(so^{(5)})/T_{\text{max}}(so^{(5)})$. Part \textbf{(b)} measures the so-called price of anarchy given by $T_{\text{max}}(uo^{(5)})/T_{\text{max}}(so^{(5)})$. This measure explains how much the travel times go up in the 5link, if users are selfish and not guided by external measures. Part \textbf{(c)} shows how much the travel times go up/down due to $E_5$ compared to the 4link system if the network is used by selfish drivers, measured by $T_{\text{max}}(uo^{(5)})/T_{\text{max}}(uo^{(4)})$. } \end{figure} In the 4link, the system optimum equals the user optimum and is thus reached by selfish drivers without any traffic guidance. In the 5link system, for selfish drivers the system optimum is only reached in the $E_5$ optimal phase (phases A$_{1/2}$). In the other phases, the system optimum would only be reached by forcing specific amounts of particles to choose specific routes via some traffic guidance system. In this case the system optimum could be reached everywhere. Figure~\ref{fig:closerlook} \textbf{(a)} shows us these potential benefits of route $E_5$. Since phase B (and B$^{\prime}$) is a "Braess 1" phase and phase D$^{\prime}$ a "$E_5$ not used-like" phase, here the system optima of 4link and 5link are equal and thus the ratio is 1. In this large phase space region the new road cannot have any positive effect - even if traffic is guided. The potential positive effect of $E_5$ is largest for low global densities and low values of $\hat{L}_{153}/\hat{L}_{14}$ (phase A$_1$) and for really high global densities (phase E$^{\prime}$). In the first case, actually all cars would use the new route and this would also be achieved without traffic guidance since in this phase $uo^{(5)}=so^{(5)}$. The ratio $T_{\text{max}}(so^{(5)})/T_{\text{max}}(so^{(4)})$ takes values as low as 1/2 meaning that the $so$ travel time can be reduced by 1/2. In phase E$^{\prime}$, the new route brings high benefits, since the 4link system here is almost full and thus travel times in the 4link rapidly diverge (see Figure~\ref{fig:t4link}~\textbf{(b)}). Here the increased capacity due to $E_5$ leads to high potential benefits, $T_{\text{max}}(so^{(5)})/T_{\text{max}}(so^{(4)})$ takes values smaller than 1/2. These full benefits would only be achieved by traffic guidance in this case, since here $uo^{(5)}\neq so^{(5)}$. In Figure~\ref{fig:closerlook} \textbf{(b)} the so-called price of anarchy~\cite{pigou2013economics} inside the 5link system given by the ratio of $T_{\text{max}}(uo^{(5)})/T_{\text{max}}(so^{(5)})$ is shown. The price of anarchy is the ratio of the maximum travel time in the user optimum (in phases A, B, C real $uo$ exist and thus all travel times are almost equal in those $uo$) and the maximum travel time in the system optimum $T_{\text{max}}(uo^{(5)})/T_{\text{max}}(so^{(5)})$. In the 4link system it is always equal to 1 since the user optimum equals the system optimum. As already shown in Figure~\ref{fig:possiblestates}, in the 5link system it is always greater or equal to 1. In phases A$_{1/2}$, the '$E_5$ optimal' phases, it is obviously equal to 1. In phase B, it is higher than 1 with values around 1.3. Not routing the traffic externally becomes really costly at high densities in the 5link system. Here the maximum travel times in the user optimum are in some cases more than 1.5 times larger than those of the system optimum. In Figure~\ref{fig:closerlook} \textbf{(c)} the ratio of the travel times in the user optima of the 4link and 5link systems $T_{\text{max}}(uo^{(5)})/T_{\text{max}}(uo^{(4)})$ is shown. This is the situation that the original Braess paradox deals with: how do the user optimum travel times in a system with selfish drivers change due to the additional road. The new road is in the case of selfish drivers only beneficial at low global densities and really high global densities. In the "Braess 1" phase (phase B) it takes values of approximately $1.2 - 1.4$ and the travel time is increased by this factor due to $E_5$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Conclusion} We analysed Braess' network with TASEP dynamics on the edges considering periodic boundary conditions and individual drivers following fixed individual strategies. Different to the case of identical particles controlled by turning probabilities at the junction sites~\citep{PhysRevE.94.062312} we could find stable travel times throughout the whole density region of both the network with and without the new edge. Without the new road, system and user optima could be found for all densities. In the major part of the phase space real user optima could also be identified in the system with the new road. Only if the new route resulting from the new road is really short compared to the old routes or the global density is really high, no real user optima exist. In the phase space region where real user optima exist, the new road leads to lower travel times only at low global densities $0 \leq \rho_{\mathrm{global}}^{(4)/(5)} \lesssim 0.2$. The largest part of the phase space is dominated by the Braess phases, with the "Braess 1" phase being the most prominent phase. The "Braess 2" phase is also found in a small phase space region. At high global densities or small path lengths of the new path no real user optima exist. This is due to the fact that here the system is prone to gridlock completely if all particles try to reduce their travel times. In this region of the phase space we chose the closest candidate of a user optimum without a possibility of gridlocks to construct the phase diagram. It turns out that here the new road also leads to higher travel times in most situations or is even completely ignored. Only at really high densities the road leads to lower travel times. Still in this whole region the system is at risk of gridlocks. This gridlock risk is also a consequence of added periodic boundary conditions and could vanish if the system was treated with open boundaries and a sufficiently large exit probability. Additionally to constructing the phase diagram we could quantify how the new road influences the system for selfish users and also what would happen if an external travel guidance authority would drive the system into its system optimum. Even if traffic was guided the network's performance, in terms of tavel times, would only be improved at really high or really low densities. The price of anarchy is relatively low inside the system with the new road. In the present system the negative implications of the new road on the network performance are even more dominant than in the version of the model at a higher level of stochasticity which we studied before~\cite{PhysRevE.94.062312}. It provides further evidence for the claim that Braess' paradox is of major importance when studying networks of microscopical traffic models. This hints at the paradox's importance in real road networks. When building new connections in a road network with the aim of reducing travel times, city planners should be very careful. Also one can see that the route choice behaviour can have enormous effects on the traffic situation. Even if on average both mechanisms in the present paper and in~\citep{PhysRevE.94.062312} are very similar, the fixed strategies lead to a different behaviour of the system and the disappearance of one of the most dominant phases observed in the system with turning probabilities. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Acknowledgements} Financial support by Deutsche Forschungsgemeinschaft (DFG) under grant SCHA~636/8-2 is gratefully acknowledged. Also support of conference travel expanses by the Bonn-Cologne Graduate School of Physics and Astronomy (BCGS) is acknowledged. Monte Carlo simulations were carried out on the CHEOPS (Cologne High Efficiency Operating Plattform for Science) cluster of the RRZK (University of Cologne). %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \clearpage \appendix \section{Measurement process} \label{sec:app-measurement-process} Here we describe how travel times in both the 4link and 5link systems were measured and how we proceeded to find user and system optima. In all our measurements, lengths were chosen as $L_0=1$, $L_1=L_3=100$, $L_2=L_4=500$ and $L_5$ and the total number of particles $M$ were varied. In all Monte Carlo simulations, random numbers were generated using the Mersenne Twister algorithm. The system was initialized randomly but particles are placed according to their strategies, i.e. randomly on the routes they want to use. A particle of strategy 153 could for example not be placed on edge $E_2$. This kind of initialization was chosen over a completely random initialization to avoid gridlocks due to wrong initialization. After initialization, the system was always relaxed for at least $5\times 10^5$ sweeps. Each particle has its own fixed strategy and for each particle and all of its finished rounds, the travel times were measured. Note that this is a different measurement process than in our previous article~\citep{PhysRevE.94.062312}. To obtain travel time values the system was kept running until the desired amount of measurements for each route was gathered. \subsection{Number of measurements} To obtain travel time values for the routes, the travel times have to be measured sufficiently often. In Figure~\ref{fig:tentwicklung} it is shown how the mean value\footnote{Note that in the main part of this article, when talking about measured travel time values, those are always the mean values of many measurements!} and standard deviation (Equation~(\ref{eq:sd}) multiplied by $\bar{T}$) of the travel times develop with the number of measurements for an example parameter set. \begin{figure}[h!] \centering \includegraphics[width=0.49\columnwidth]{./4link-tentwicklung-L5-278-M-148.pdf} \includegraphics[width=0.49\columnwidth]{./5link-tentwicklung-L5-278-M-148.pdf} \caption{\label{fig:tentwicklung} The development of the mean $\bar{T}$ and the standard deviation $\sigma(\bar{T})$ of the travel time values against the number of sweeps. The values for routes 14/23/153 are shown in \textcolor{purple-density}{purple}/\textcolor{brown-density}{brown}/\textcolor{blue-density}{blue}. Part \textbf{(a)} shows measurements of the 4link system for $M=148$ and $n_{\text{l}}^{(j_1)}=0.5$ and part \textbf{(b)} shows measurements of the 5link system with $L_5=278$ and $M=148$ and $n_{\text{l}}^{(j_1)}\approx 0.92$ and $n_{\text{l}}^{(j_2)}\approx 0.10$. The second y-axes show how many measurements $n_{14/23/153}$ of $T_{14/23/153}$ were performed. The curve with the long dashes represent the number of measurements of the travel times of route 14 and 23, the curve with the smaller dashes the number of measurements of the travel time of route 153.} \end{figure} \clearpage \subsection{Sweeping the network state - landscapes to find user and system optima} \label{sec:app-sweep} A simple yet CPU time costly way to find the user and system optima is to check all different combinations of $N_{14}$, $N_{23}$ and $N_{153}$, or all combinations $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$, both varying from 0 to 1, for a given $(L_5, M)$ and see, which combinations minimize $\Delta T$ and $T_{\text{max}}$ respectively. We used a grid resolution of 0.1 for finding the system optima of the 5link system. An example of how the $\Delta T$ and $T_{\text{max}}$ landscapes look like can be seen in Figure~\ref{fig:sweep-landscape}. It should be noted that the actual system optima may in most cases lie in between the 0.1 grid points. Nevertheless, this grid is sufficient to see whether the system optimum of the 5link system has a lower or higher maximum travel time than that of the 4link system for a given $(L_5, M)$. It has to be noted though that this technique does not ensure that we found the actual system optima, which is why the coloured points in Figures~\ref{fig:closerlook} \textbf{(a)} and \textbf{(b)} should not be seen as accurate values. For finding user optima we used a more accurate method as described in~\ref{sec:MC}. \begin{figure}[h!] \centering \includegraphics[width=0.49\columnwidth]{./traveltimes-L0-1-L13-100-L24-500-L5-278-M-148-so.pdf} \includegraphics[width=0.49\columnwidth]{./traveltimes-L0-1-L13-100-L24-500-L5-278-M-148-uo.pdf} \caption{\label{fig:sweep-landscape}The $\Delta T$ and $T_{\text{max}}$ landscapes of the 5link system for $L_5=278$ and $M=148$ depending of the $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$. As can be seen, this is an "$E_5$ optimal" state, since the minima of $\Delta T$ and $T_{\text{max}}$ coincide at $n_{\text{l}}^{(j_1)}=0.9$ and $n_{\text{l}}^{(j_2)}=0.1$ and the corresponding strategy has a lower maximum travel time than the 4link's system optimum which is found at $n_{\text{l}}^{(j_1)}=0.5$ and $n_{\text{l}}^{(j_2)}=1.0$.} \end{figure} \subsection{Metropolis algorithm for finding user optima} \label{sec:MC} The method of sweeping the $\Delta T$ and $T_{\text{max}}$ landscapes to find user optima, as described in~\ref{sec:app-sweep}, is not very effective. If the whole region $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$ both from 0 to 1 is sweeped, a lot of measurements are performed far away from the user (and system) optima. Those measurements are a waste of CPU time. Furthermore, depending on how fine the step size is, the user optima will most likely not lie on one of the scanned grid points but in between. A brute force way would be rescanning the corresponding landscape regions with a finer grid. Since this is very time intensive we developed a Metropolis Monte-Carlo~\cite{doi:10.1063/1.1699114} algorithm for finding user optima. The algorithm works like this: \begin{enumerate} \item Set maximum step width $sw$ and 'temperature' $T$ \item Set start values $(n_{\text{l}}^{(j_1)}, n_{\text{l}}^{(j_2)})$ and from this \begin{itemize} \item $N_{14}=M\cdot n_{\text{l}}^{(j_1)}\cdot n_{\text{l}}^{(j_2)}$ \item $N_{23}=M\cdot(1-n_{\text{l}}^{(j_1)})$ \item $N_{153}=M\cdot n_{\text{l}}^{(j_1)}\cdot (1-n_{\text{l}}^{(j_2)})$ \end{itemize} \item Let the system thermalize with strategy according to $(n_{\text{l}}^{(j_1)}, n_{\text{l}}^{(j_2)})$ \item Measure travel times $T_{14}$, $T_{23}$, $T_{153}$ and calculate $\Delta T$ \item Suggest new $(n_{\text{l, new}}^{(j_1)}, n_{\text{l,new}}^{(j_2)})$ by drawing a random number $z$ between 0 and 2$\pi$ and setting $(n_{\text{l, new}}^{(j_1)}, n_{\text{l, new}}^{(j_2)})=(n_{\text{l}}^{(j_1)}+sw\cdot \cos(z), n_{\text{l}}^{(j_2)}+sw\cdot \sin(z))$ and calculate $N^{\text{new}}_{14}$, $N^{\text{new}}_{23}$, $N^{\text{new}}_{153}$ as in step 2 \item Let the system thermalize with strategy according to $(n_{\text{l, new}}^{(j_1)}, n_{\text{l, new}}^{(j_2)})$ \item Measure travel times $T^{\text{new}}_{14}$, $T^{\text{new}}_{23}$, $T^{\text{new}}_{153}$ and calculate $\Delta T^{\text{new}}$ \item Accept the new strategy with probability $p=\min\left(1, \exp\left(-\frac{\Delta T-\Delta T^{\text{new}}}{T}\right)\right)$ \item Repeat steps 5 to 7 as long as $\Delta T^{\text{new}}>\epsilon$, with tolerance $\epsilon$ \end{enumerate} In this algorithm, the maximum step width $sw$ is the maximum possible value, $n_{\text{l}}^{(j_1)}$ and $n_{\text{l}}^{(j_2)}$ can be changed by. The temperature $T$ is a measure for the probability with which a strategy with higher $\Delta T$ might be accepted and $\epsilon$ is the tolerance: if $\Delta T \leq \epsilon$ the strategy is accepted as the user optimum. The 'real' user optimum is reached, if $\epsilon$ is exactly zero. It turned out to be useful to set $\epsilon=20$, $sw=0.1$ and $T=10$. An additional tenth step could be added to the algorithm, in which $sw$ would be reduced, if newly suggested probabilities get rejected a certain amount of times. This additional step was not needed in our case. The algorithm performed well in most of our situations. Fig.~\ref{fig:mc-verlauf}~\textbf{(a)} shows the search path of the algorithm for $L_5=278$ and $M=148$ for 10 different start values $(n_{\text{l}}^{(j_1)}, n_{\text{l}}^{(j_2)})$. The $uo$-landscape with 0.1 step width as described in the previous subsection is underlayed for visualization purposes. \begin{figure}[h!] \centering \includegraphics[width=0.49\columnwidth]{./mc-verlauf-landscape.pdf} \includegraphics[width=0.49\columnwidth]{./mc-verlauf.pdf} \caption{\label{fig:mc-verlauf}An example of the performance of the Metropolis algorithm for $L_5=278$ and $M=148$ and ten different start values. In \textbf{(a)} the search paths are shown with underlayed values of $\Delta T$ which were obtained by sweeping the $(n_{\text{l}}^{(j_1)}, n_{\text{l}}^{(j_2)})$-landscape as described in~\ref{sec:app-sweep}. The beginnings of all paths are marked by a $\bigcirc$ and the endings by a $\bigtriangleup$. Also the user optimum is marked by a white $\times$. In \textbf{(b)} the corresponding $\Delta T$ values against the number of Metropolis steps are shown. One can see, that the algorithm converges pretty fast for all 10 start values.} \end{figure} Furthermore, in Fig.~\ref{fig:mc-verlauf}~\textbf{(b)} the $\Delta T$ values against the Metropolis step nummber (i.e. how often steps 5 to 8 of the algorithm were performed) is shown. From both pictures it can be deduced that the algorithm works really well for this case. Indeed, it performed really well for most other $(L_5, M)$ combinations as well. Sometimes, depending on the start values, the algorithm will not converge and has to be restarted with different start values. Also in some cases there is more than one user optimum, as described in~\ref{sec:app-special}. In this case, the algorithm will get 'caught' in one of them. The algorithm can also be used to find system optima if after each step $T_{\text{max}}$ is calculated and the newly suggested strategy is accepted if $T_{\text{max}}$ got lower. The problem in this case is that there is no real termination condition as there is no a priori known lower bound to $T_{\text{max}}$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Special cases} \label{sec:app-special} For some configurations we could find more than one user optimum. Here we present the example of $\hat{L}_{153}/\hat{L}_{14}=0.4$ and $\rho^{(5)}_{\text{global}}=0.18$. In Figure~\ref{fig:specialcases} the $T_{\mathrm{max}}$ and $\Delta T$ landscapes for this parameter set are shown. In the figure, the values from a sweep of the landscapes in steps of 0.1 is underlayed. \begin{figure}[h] \centering \includegraphics[width=0.49\columnwidth]{./traveltimes-L0-1-L13-100-L24-500-L5-37-M-224-so.pdf} \includegraphics[width=0.49\columnwidth]{./traveltimes-L0-1-L13-100-L24-500-L5-37-M-224-uo.pdf} \caption{\label{fig:specialcases}The $T_{\mathrm{max}}$ (part \textbf{(a)}) and $\Delta T$ (part \textbf{(b)}) landscapes of the 5link system with $L_1=L_3=100$, $L_2=L_4=500$, $L_5=37$, $M=224$, $\rho^{(5)}_{\text{global}}=0.18$. In part \textbf{(a)}, the system optimum is marked by $+ 1$. In part \textbf{(b)} one can see that there are three different user optima, $\times 1$ to $\times 3$, the values of which are given in Table~\ref{tab:special}.} \end{figure} The travel time and $\Delta T$ and $T_{\mathrm{max}}$ values for the four marked points are given in Table~\ref{tab:special}. From part \textbf{(a)} of the picture one can see that the system optimum is given by $n_{\text{l}}^{(j_1)}=0.5$ and $n_{\text{l}}^{(j_2)}=1.0$. This means that here $so^{(5)}=so^{(4)}$ since this is the state where half of the particles use route 14 and the other half route 23. In part \textbf{(b)} of the Figure one can see that there are three different user optima. The two user optima $\times 1$ and $\times 3$ were found by sweeping the $n_{\text{l}}^{j_1/j_2}$ and are special cases of user optima which were already mentioned in Section~\ref{sec:mc-observables}. The user optimum $\times 2$ was found by our Metropolis algorithm. The optimum $\times 1$ is a special case since only paths 23 and 153 are used. Since both their travel times are almost equal and smaller than that of the unused route 14 this state is a user optimum. For calculating $\Delta T$, only the difference between $T_{23}$ and $T_{153}$ is used. It would in this case not make sense for any particle to switch to route 14 which has a higher travel time. The same happens in the user optimum $\times 3$, but here routes 14 and 153 are used and route 23 is not. The other user optimum $\times 2$ is an 'ordinary' user optimum in which all three routes are used and have (almost) the same travel time. \begin{table}[h] \caption{\label{tab:special} The $n_{\text{l}}^{(j_1/j_2)}$, $\Delta T$, $T_{\mathrm{max}}$ and travel time values of the routes for the 4 points which are marked in Figure~\ref{fig:specialcases}. The points $\times 1$ to $\times 3$ are three user optima of the system, while $+ 1$ is the system optimum.} \begin{center} \begin{tabular}{| c | l | l | l | l | l | l | l |} \hline Point & $n_{\text{l}}^{(j_1)}$ & $n_{\text{l}}^{(j_2)}$ & $T_{14}$ & $T_{23}$ & $T_{153}$ & $\Delta T$ & $T_{\mathrm{max}}$ \\ \hline $\times 1$ & 0.5 & 0.0 & 926 & 880 & 876 & 4 & 880 \\ \hline $\times 2$ & 0.808 & 0.221 & 970 & 975 & 975 & 10 & 975 \\ \hline $\times 3$ & 1.0 & 0.5 & 878 & 1136 & 875 & 3 & 878 \\ \hline\hline $+ 1$ & 0.5 & 1.0 & 743 & 742 & 294 & 898 & 743 \\ \hline \end{tabular} \end{center} \end{table} The (maximum) travel time in all three user optima is higher than that of the system optimum (which is the same as the 4link system optimum) which leads to the conclusion that no matter in which user optimum the system ends up, a "Braess 1" state is present. While in the whole unhatched area of the phase diagram (Figure~\ref{fig:phases}) user optima were found we cannot guarantee that \textit{all} user optima were found. Also, the values (given by the color of the points) in Figures~\ref{fig:closerlook} \textbf{(a)} to \textbf{(c)} are given for one of the found user optima and could be slightly different if, for the cases of multiple user optima, another user optimum was chosen. This is because the user optima, as in the example presented here (see Table~\ref{tab:special}), could have different travel time values. Therefore the values in Figure~\ref{fig:closerlook} should not be interpreted as exact values but as a tendency of how the system changes due to the addition of $E_5$. The fact that we found multiple user optima with different travel times (and different $T_{\mathrm{max}}$ values and also different total travel time values) for the same parameter set is also a difference to what is observed in mathematical models of road traffic. In these models it was shown that "$[$the user optimum$]$ is unique whenever the shortest routes between all pairs of locations are unique and cost is strictly increasing with increasing flow"~\cite{beckmann1956studies}. \clearpage %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{References} \bibliography{paper-braess-2} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \end{document} }
\caption{\textbf{Opportunity to find the exemplar VM instances across workloads for reducing operational cost.} The \emph{y-axis} represents the percentage of workloads (out of 107 in three systems) that are within 30\% difference with the optimal performance. The colored bars are VM types that considered the exemplar configurations for the majority of workloads ($>= 50\%$). The \textcolor{red}{red} bar represents that the VM type is more likely to be the optimal choice.}
\caption{\small (a) Examples of AU predictions: ground-truth (top), fusion module (middle) and structure inference (bottom) prediction (\textcolor{green}{$\bullet$}: true positive, \textcolor{red}{$\bullet$}: false positive). (b) AUs correlation in BP4D (\textcolor{green}{$\bullet$}: positive, \textcolor{red}{$\bullet$}: negative). Line thickness is proportional with correlation magnitude. (c) Class activation map for AU24 that shows the discriminative regions of simple patch prediction (left) and DSIN (right). Best seen in color.}
\caption{ Residuals of the best fit diffuse emission model to the AGILE map. The white circle shows the ring of \gray-photon excess centred on $\kappa$ Orionis (white star). The green dot on the top-left part of the map highlights the excess of emission corresponding to the unidentified EGRET source 3EG J0616-0720. }
\caption{ Left panel: the ring of \gray excess detected by AGILE. Right panel: the CO map from \cite{Dame01}, which reveals the star forming shell discussed by \cite{Pillitteri16}; the contour levels from the \gray data are shown in black. }
\caption{\small Formation enthalpies of the {\bf (a)} TlBiS$_2$-TlBiSe$_2$ {\bf (b)} TlBiSe$_2$-TlBiTe$_2$ {\bf (c)}TlBiS$_2$-TlBiTe$_2$ structures using DFT calculations (\textcolor{pranab_red}{$\bullet$}) and cluster expansion technique (\textcolor{pranab_green}{$\times$}) (upper panel). Their respective binodal ({\color{orange}{---}}) and spinodal curves ({\color{brown}{- - -}}) are illustrated in the lower panels {\bf (d,e,f)}. {The eutectic isotherm, from Ref.~\cite{Jafarov_im_2014,Babanly_amchemscij_2016} is illustrated with} ({\color{pranab_green}{$\cdot\cdot\cdot$}}). }
\caption{\small Electronic band structure of (TlBiSe$_2$)$_{1-x}$(TlBiTe$_2$)$_x$ boundaries for compositions {\bf (a)} $x={2}/{18}$, {\bf (b)} $x={3}/{18}$, {\bf (c)} $x={14}/{18}$, {\bf (d)} $x={15}/{18}$, and {\bf (e)} $x={16}/{18}$. The contribution of four atomic layers from each side of the region, or the contribution of the entire thin region (in the case of very thin regions), are projected onto the electronic band structures. {\bf (f)} and {\bf (h)} Projected and isosurface representation of {summed} charge density for $x={2}/{18}$, ${3}/{18}$, and ${14}/{18}$. {\bf (g)} Schematic representation of band alignment of (TlBiSe$_2$)$_{1-x}$(TlBiTe$_2$)$_x$ boundaries for TlBiSe$_2$-rich and TlBiTe$_2$-rich compositions. The insets show the band structure with in-plane spin contributions. The size of symbols are proportional to the degree of in-plane spin polarization in the five atomic layers nearest the interface, and the red and blue colors correspond to the in-plane spin-up and spin-down directions, respectively (color code: \textcolor{pranab_red}{$\bullet$}Bi~~~\textcolor{brown}{$\bullet$}Tl~~~\textcolor{green}{$\bullet$}Se~~~\textcolor{PaleBlue}{$\bullet$}Te). }
\caption{\small Electronic band structure of (TlBiS$_2$)$_{1-x}$(TlBiTe$_2$)$_x$ boundaries for compositions {\bf (a)} $x={1}/{18}$, {\bf (b)} $x={2}/{18}$, {\bf (c)} $x={3}/{18}$, {\bf (d)} $x={13}/{18}$, {\bf (e)} $x={14}/{18}$, and {\bf (f)} $x={15}/{18}$. The contribution of four atomic layers from each side of the region, or the contribution of the entire thin region (in the case of very thin regions), are projected onto the electronic band structures. {\bf (g)} Projected and isosurface representation of {summed} charge density {at the Dirac point} for $x={1}/{18}$, $x={2}/{18}$ and $x={14}/{18}$. The insets show the band structure with in-plane spin contributions. The size of symbols are proportional to the degree of in-plane spin polarization in the five atomic layers nearest the interface, and the red and blue colors correspond to the in-plane spin-up and spin-down directions, respectively (color code: \textcolor{pranab_red}{$\bullet$}Bi~~~\textcolor{brown}{$\bullet$}Tl~~~\textcolor{pranab_green}{$\bullet$}S~~~\textcolor{PaleBlue}{$\bullet$}Te). }
\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.}
\caption{Comparison of a deep neural network with linear models. Models are trained to predict the direction $\{-1, +1\}$ of next mid-price move. Comparison for approximately $500$ stocks and out-of-sample results reported for June-August, 2015. Left-hand figure: increase in accuracy of stock-specific deep neural networks versus stock-specific linear models. Right-hand figure: accuracy of a universal deep neural network (\textcolor{red}{red}) compared to stock-specific linear models (\textcolor{blue}{blue}).}
\caption{Performance on $500$ new stocks which the model has never seen before. \textcolor{blue}{Left}: out-of-sample accuracy reported for June-August, 2015. \textcolor{red}{Right}: out-of-sample accuracy reported for January-March, 2017. Universal model trained on data from January 2014-May 2015.}
\caption{Out-of-sample increase in accuracy when using \textcolor{red}{5000} lags versus \textcolor{blue}{100} lags, across $1,000$ stocks. Test period : June-August 2015.}
\caption{\textbf{Magnetic WSM order in equilibrium} \textbf{a}, Phase diagram as a function of Hubbard $U$ and hopping integral, $t_\sigma$, at \blue{$T=0.016$ eV (186 K)}. The red dashed line indicates the chosen parameter subspace for our calculations at $t_\sigma = -0.62$ $\mathrm{eV}$. For small interactions, $U \leq \blue{1.22}$ $\mathrm{eV}$, the system exhibits a paramagnetic metallic phase (PMM). In the intermediate regime, \blue{$1.22$ $\mathrm{eV}$ $< U \leq 1.25$ $\mathrm{eV}$}, TRS is broken by spontaneous magnetic all-in/all-out (AIAO) order. A further increase of $U$ above \blue{$1.25$ $\mathrm{eV}$} leads to a gapped antiferromagnetic insulating phase (AFI) with larger magnetization. For $|t_\sigma|\geq 0.62$ $\mathrm{eV}$ the WSM-AFI transition is of 2nd order, indicated by black dots. Below that value, the transition is of 1st order (bicoloured dots). The dots' increasing size corresponds to the increasing region of hysteresis. The red square indicates the chosen initial state ($U_{i} = 1.28$ $\mathrm{eV}$) in parameter space. The red arrow shows the intended quenching direction. \textbf{b}, Magnetization as a function of $U$ for the choice of $t_\sigma = -0.62$ $\mathrm{eV}$. The dashed lines indicate the region of hysteresis. The coloured arrows show the final values of $U=U_q$ right after the interaction quench for the non-equilibrium calculations (see FIG. \ref{fig:2}).}
\caption{\textbf{Optically induced nonthermal Weyl semimetal.} \textbf{a}, Self-consistent TDDFT+U calculation of $U_{\textit{eff}}(t)$ (Ir 5d) for $\mathrm{Y_2Ir_2O_7}$ under pump excitation ($0.41$ eV frequency, $25.4$ fs duration). $U_{\textit{eff}}$ decreases, initially following the pulse envelope. After 20 fs it reaches a plateau, which decreases for higher intensities. Relative changes of $U_{\textit{eff}}$ are $1.2\%$ (blue), $5.3\%$ (green), $9.6\%$ (red) and $40.4\%$ (cyan). \textbf{b}, Time-dependent AIAO magnetization from TDDFT+U. For the smallest intensity (blue) the magnetization drops by $25\%$. For all higher pump intensities, there is a massive decrease of the magnetization by $88\%$ (red) and $92\%$ (green, brown), respectively. \textbf{c}, Time-dependent AIAO magnetization $m(t)$ (model) \blue{for bath coupling $\Gamma_0 = 0.008$ eV} after instantaneous quenches at $t=0$ from $U_{i} = 1.28$ eV to $U_q = 1.24$ eV, $1.20$ eV and $1.16$ eV, respectively \blue{(dashed curves are without bath, $\Gamma_0 = 0$)}. Thermal $m$ values for $U_q$ are indicated by coloured arrows for reference. \blue{After a fast drop followed by an over-damped oscillation, the magnetization slowly converges towards the respective thermal value.} \textbf{d}, Equilibrium phase diagram as a function of $T$ and $U$ with optically induced nonthermal states. Effective temperatures $T_{\textit{eff}}$ (see Supplementary Fig.~\ref{fig:suppenergy}) of the nonequilibrium states \blue{at $t=123.4$ fs} after the quench from the initial thermal state (red square), for quench values $U_q$ as indicated (see Supplementary Fig.~\ref{fig:supptime}). \blue{Pentagons (squares) indicate nonequilibrium WSM states for the open system (closed system).} (see Figure \ref{fig:3}). The black diamond indicates a nonequilibrium AFI state. }
\caption{\textbf{Time-resolved ARPES detection of Weyl fermions.} \textbf{a}, Calculated photocurrent from tr-ARPES of the equilibrium band structure along high-symmetry path for an initial AFI state ($U_{i}=1.28$ $\mathrm{eV}$, $t_\sigma = -0.62$ $\mathrm{eV}$) at \blue{$T=0.016$ eV}. The conduction bands are separated from the valence bands by an energy gap $E_G \approx 0.15 $ $\mathrm{eV}$ (AFI) at the $L$ point. \blue{The yellow square highlights the region of interest.} \textbf{b}-\textbf{d}, tr-ARPES band structure along the high-symmetry line $\mathrm{L - \Gamma}$ at probe time \blue{$t_{p} = 123.4 $ $\mathrm{fs}$}. The quenched interaction values, $U_q = 1.24$ $\mathrm{eV}$, $1.20$ $\mathrm{eV}$ and $1.16$ $\mathrm{eV}$, correspond to increasing $\Delta U$ between initial and quenches $U$ and thus stronger driving fields. The black solid lines show the instantaneous band structures of the Hamiltonian at $t=t_{p}$. The coloured arrows indicate the dynamically generated Weyl points. The Weyl cones are shifted along the high symmetry line towards $\mathrm{\Gamma}$-point with increasing $\Delta U$.}
\caption{\textbf{Fate of nonthermal WSM for relaxing $U(t)$.} \textbf{a}, Time-dependent interaction relaxing back to the initial value, \blue{${U(0 \le t \le 150 \mathrm{fs}) = U_i-(U_i - U_q)\sin{\left(\pi\cdot t/150 \mathrm{fs}\right)}}$}. \textbf{b}, Corresponding time evolution of the magnetization. As in the instantaneous case, the magnetization is initially reduced before slowly relaxing back towards its initial value. The vertical solid lines indicate the probe times $t_p$ for which the tr-ARPES signals are shown in \textbf{c}-\textbf{h}, as indicated, with a probe width \blue{$\sigma_{p} = 20.6$ $\mathrm{fs}$} shown by the grey-shaded area for one case. \textbf{c}-\textbf{h}, Tr-ARPES signals at different probe times as indicated along $\mathrm{L-\Gamma}$. The solid lines correspond to the effective band structures of the time-dependent Hamiltonian averaged over the FWHM of the probe pulse. }
\caption{\textbf{Nonequilibrium magnetization and energy.} \textbf{a}, \blue{Closed system ($\Gamma_0 = 0$)} time evolution of the magnetic order parameter for different quench values $U_q$, varied in steps of $0.02$ $\mathrm{eV}$ within the interval $\left[1.12, 1.26\right]$ $\mathrm{eV}$. \textbf{b}, Nonequilibrium energy per particle per unit cell for the corresponding interaction interval. The system energy only changes at t=0, at which the (closed) system is quenched. The amount of energy pumped into the system scales linearly with $\Delta U$. It remains constant before and after the quench. The final energy values are used to calculate the effective nonequilibrium temperatures $T_{\textit{eff}}$, displayed in Figure \ref{fig:2}(c). \blue{\textbf{c}, Open system ($\Gamma_0 = 0.008$ eV) magnetization dynamics of the open system for same parameter set. \textbf{d}, Time-dependent total energy of the open system. After the quench at $t=0$, the total energy relaxes towards its thermal steady-state value on a time scale $\Gamma_0^{-1} \approx 80$ fs. The dashed vertical line denotes the probe time, $t_{p} = 123.4 $ $\mathrm{fs}$, used for Figure \ref{fig:3} (b-d).}} \label{fig:supptime} \end{figure} %%%%% FIG.2c SUP1 %%%%% \begin{figure}[htp!] \centering \includegraphics[width=1.0\linewidth]{FIG2_d_SUP_1.pdf} \caption{\textbf{Temperature dependence of total energy and magnetization.} \blue{\textbf{a}, Temperature-dependent equilibrium mean energy per particle per unit cell for different values $U$, varied in steps of $0.02$ $\mathrm{eV}$ within the interval $\left[1.12, 1.28\right]$ $\mathrm{eV}$. From these curves, the effective temperatures $T_{\textit{eff}}$, assigned to the nonequilibrium states in Figure \ref{fig:supptime}(b, d), can be read off by attributing each energy value at the probe time, $t_p = 123.4$ fs, to its corresponding temperature. \textbf{b}, $\Delta \equiv U\cdot m$ in dependence of the equilibrium temperature. The WSM-PMM phase boundary can be read off at that point where $\Delta(T)$ reaches zero. $\Delta_{crit} \approx 0.24$ $\mathrm{eV}$ denotes the critical value at which the AFI-WSM transition takes place. The dip appearing for intermediate interactions is associated with the semi-metallic band structure and change in density of states in the WSM phase as opposed to the insulating AFI phase.}
\caption{Performance comparison with the state-of-the-art VQA methods on the test-standard split of VQA v2 dataset. \colorbox[rgb]{ .682, .667, .667}{BUTD-ensemble} is an ensemble of 30 models and it will not participate in ranking. Accuracies in percentage (\%) are reported.}
\caption{{\color{red}\bf Please uncouple the panels. There are no space considerations in the SM, so we should not include multiple graphs in a panel. Please send me each of these graphs as a separate file, labeled figure 1a, 1b, etc.\\ The main text should have the same figure as the left panel here. Is this the one with the ln2 term included? \\ Please use another, other than grey, color for the dashed curves. Red dashed curves might look easier to see. }\\ Left panel: GP prediction (solid curves) of the free energy density $f(T,\tilde{m})$ of the mean-field Heisenberg model produced by Eq. (\ref{eqn:f}) (dashed curves). Inset: the order parameter m0 that minimizes $f(T,\tilde{m})$: symbols -- GP predictions, dashed curve -- Eq. (\ref{eqn:f}). Right panel: GP prediction (solid curves) of $f(T,\tilde{m})$ for $T \geq T_c$ for the constructed kernels at each iteration of the proposed algorithm. The markers illustrate order parameter $\tilde{m}_0$ which minimizes the predicted $f(T,\tilde{m})$ with GP regression. The bottom inset illustrates the change of $\tilde{m}_0$ for each proposed kernel combination as a function of $T$. All GP models are trained with 330 points at $1.47 < T < 2.08$ (shaded area) and $-1.25 < \tilde{m} < 1.25$. \label{fig:H_f1} }
\caption{Comparison of ATE RMSE (in cm) with alternative and baseline methods on TUM sequences~\protect\cite{sturm12iros}. Colors indicate the {\color{blue}best} and {\color{green}second best} results.}
\caption{Qualitative evaluation on 6 video sequences (\ie \emph{CarScale}, \emph{Dog}, \emph{Girl2}, \emph{Human3}, \emph{Panda} and \emph{Trans}). We show the results of {\color{green}STRCF}, {\color{yellow}ECO-HC}, {\color{blue}BACF}, {\color{red}SRDCF} and {\color{cyan}SRDCFDecon} with different colors, respectively.}
\caption{Second derivative plots of the ARPES spectra for (a) clean Au(111) and (b) (4x4)-BlueP on Au(111) along the $\Gamma \rm K$, respectively, using horizontally polarized light with h$\nu = 60$ eV. Second derivative plots of the ARPES spectra for (c) clean Au(111) and (d) BlueP on Au(111) along the $\Gamma \rm M$, respectively, using horizontally polarized light with of h$\nu = 55$ eV. (e), (f) Close-up in the energy region from Fermi level down to 2 eV BE for the spectra of (b) and (d), respectively. (g) Energy distribution curves at \kpara\= 0.8\invA\for the BlueP on Au(111) along the\GbarKbarBlueP\high-symmetry direction for three different excitation energies. (h) Momentum distribution curves extracted from (b) at 0.95 and 1.25 eV BE, marked by red and blue dashed lines in (b), respectively. (i) Red and green lines correspond to energy distribution curves at\kpara\= 0.95\invA\and\kpara\= 1.31\invA\along the\GbarKbarBlueP\direction, respectively, extracted from (b). The black line corresponds to the energy distribution curve along the\GbarMbarBlueP\high-symmetry direction of (d) at\kpara\= 0.92\invA. The momentum positions of the extracted energy distribution curves are marked with the corresponding dashed line color in (b) and (d). (j) Effective band structure of (4x4)-BlueP on Au(111) as presented in Fig. \ref{fig2}(f) unfolded to the primitive cell of Au(111) surface. In the subfigures, gold and blue letters mark the high-symmetry points of Au(111) and freestanding BlueP, respectively. Distances of high-symmetry segments in the reciprocal space for BlueP are \GbarKbarBlueP=1.26 \invA, \GbarMbarBlueP=1.09 \invA, and for Au(111) \GbarKbar=1.44 \invA, and \GbarMbar=1.25 \invA, respectively.}
\caption{Constant energy contours of (4x4)-BlueP on Au(111) at BEs (a) 0.025 eV (b) 0.85 eV and (c) 1.165 eV. Blue bashed lines in (a) border part of the first BZ of BlueP. (d) Band dispersions along the high-symmetry directions marked with blue dashed lines in (a). (e), (f) Second derivative plots of the ARPES spectra along the \GbarKbarBlueP, \GbarMbarBlueP\directions, respectively. All ARPES spectra were acquired using horizontally polarized light with h$\nu = 60$ eV. (g) Momentum distribution curves extracted from (e), ((f)) at BEs of 0.92 eV using blue dashed line (green dashed line) and 1 eV red dashed line (black dashed line), respectively. (h) Energy distribution curves at the momentum positions marked in (d).}
\caption{Occurrence of the edges between hateful \tikz{\definecolor{tempcolor}{RGB}{248,65,74} \draw[fill=tempcolor,line width=1pt] circle(0.8ex);} and normal \tikz{\definecolor{tempcolor}{RGB}{86,118,161} \draw[fill=tempcolor,line width=.5pt] circle(0.8ex);} users, and between suspended \tikz{\definecolor{tempcolor}{RGB}{255,250,205} \draw[fill=tempcolor,line width=1pt] circle(0.8ex);} and active \tikz{\definecolor{tempcolor}{RGB}{239,204,0} \draw[fill=tempcolor,line width=.5pt] circle(0.8ex);} users. Results are normalized w.r.t. to the type of the source node, as in: $P($source type$\rightarrow$dest type$|$source type$)$. Notice that the probabilities do not add to $1$ in hateful and normal users as we don't present the statistics for non-annotated users.}
\caption{2NN\label{fig-x2NN}}{\includegraphics[width=\textwidth]{percol_4D-pics-10-pdfjam}}
\caption{2NN+NN\label{fig-x2NN_NN}}{\includegraphics[width=\textwidth]{percol_4D-pics-11-pdfjam}}
\caption{3NN\label{fig-x3NN}}{{\includegraphics[width=\textwidth]{percol_4D-pics-12-pdfjam}}}
\caption{3NN+NN\label{fig-x3NN_NN}}{{\includegraphics[width=\textwidth]{percol_4D-pics-13-pdfjam}}}
\caption{3NN+2NN\label{fig-x3NN_2NN}}{{\includegraphics[width=\textwidth]{percol_4D-pics-14-pdfjam}}}
\caption{3NN+2NN+NN\label{fig-x3NN_2NN_NN}}{{\includegraphics[width=\textwidth]{percol_4D-pics-15-pdfjam}}}
\caption{ Unsupervised re-id performance evaluation. % with state-of-the-art methods. {\bf Metric}: Rank-1 and mAP (\%). The $1^\text{st}/2^\text{nd}$ best results are in {\color{red}red} and {\color{blue}blue}. TJ-AIDL$^\text{Duke}$ / TJ-AIDL$^\text{Market}$: Our TJ-AIDL using DukeMCMT-ReID and Market-1501 as the labelled source, respectively. %domain with labelled identity and attributes. }
\caption{\footnotesize Comparing bias and MSE (mean-squared error) of estimators across different left and right truncation proportions, under \textit{dependent} survival and truncation times. Survival times generated from proportional hazards model with hazard function $\lambda(t)\exp(\beta_1Z_1 + \beta_2Z_2)$, with $\beta_1=\beta_2=1$. Survival times conditionally independent of left and right truncation times given $Z_1$. For $j=1$ (left column) and $j=2$ (right column): Top row compares bias of proposed EM estimator $\widehat{\beta}_{em,j}$ (\textbf{black}) to weighted estimator $\widehat{\beta}_{w,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which does not account for dependent truncation. Middle row compares bias of $\widehat{\beta}_{em,j}$ (\textbf{black}) to the standard estimator $\widehat{\beta}_{s,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which ignores truncation completely. Bottom row compares rMSE($\widehat{\beta}_{em,j})$ (\textbf{black}) to rMSE($\widehat{\beta}_{w,j})$ (\textcolor{dark-gray}{\textbf{gray}}). Here rMSE($\widehat{\beta}$) = $\frac{\text{MSE}(\widehat{\beta}_j)}{\text{MSE}(\widehat{\beta}_{s,j})}$ is the relative MSE of the estimator $\widehat{\beta}_j$ to the standard estimator $\widehat{\beta}_{s,j}$.}
\caption{\footnotesize Comparing bias and MSE (mean-squared error) of estimators across different left and right truncation proportions, under \textit{independent} survival and truncation times. Survival times generated from proportional hazards model with hazard function $\lambda(t)\exp(\beta_1Z_1 + \beta_2Z_2)$, with $\beta_1=\beta_2=1$. For $j=1$ (left column) and $j=2$ (right column): Top row compares bias of proposed EM estimator $\widehat{\beta}_{em,j}$ (\textbf{black}) to weighted estimator $\widehat{\beta}_{w,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which does not account for dependent truncation. Middle row compares bias of $\widehat{\beta}_{em,j}$ (\textbf{black}) to the standard estimator $\widehat{\beta}_{s,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which ignores truncation completely. Bottom row compares rMSE($\widehat{\beta}_{em,j})$ (\textbf{black}) to rMSE($\widehat{\beta}_{w,j})$ (\textcolor{dark-gray}{\textbf{gray}}). Here rMSE($\widehat{\beta}$) = $\frac{\text{MSE}(\widehat{\beta}_j)}{\text{MSE}(\widehat{\beta}_{s,j})}$ is the relative MSE of the estimator $\widehat{\beta}_j$ to the standard estimator $\widehat{\beta}_{s,j}$.}
\caption{\footnotesize Comparing bias and MSE (mean-squared error) of estimators under \textit{dependent left truncation}. Here $\rho_{LT}$ is the correlation between the left truncation and survival time. Survival times generated from proportional hazards model with hazard function $\lambda(t)\exp(\beta_1Z_1 + \beta_2Z_2)$, with $\beta_1=\beta_2=1$. For $j=1$ (left column) and $j=2$ (right column): Top row compares bias of proposed EM estimator $\widehat{\beta}_{em,j}$ (\textbf{black}) to weighted estimator $\widehat{\beta}_{w,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which does not account for dependent truncation. Middle row compares bias of $\widehat{\beta}_{em,j}$ (\textbf{black}) to the standard estimator under left truncation $\widehat{\beta}_{s_L,j}$ (\textcolor{dark-gray}{\textbf{gray}}), which accounts for dependent left truncation. Bottom row compares rMSE($\widehat{\beta}_{em,j})$ (\textbf{black}) to rMSE($\widehat{\beta}_{w,j})$ (\textcolor{dark-gray}{\textbf{gray}}). Here rMSE($\widehat{\beta}$) = $\frac{\text{MSE}(\widehat{\beta}_j)}{\text{MSE}(\widehat{\beta}_{s_L,j})}$ is the relative MSE of the estimator $\widehat{\beta}_j$ to the standard estimator under left truncation $\widehat{\beta}_{s_L,j}$.}
\caption{Overview of the agent's decision making algorithm. Methods that take place in other objects, such as applying actions to the world or selecting a random number are omitted for brevity. HORIZON is $n+1$ for $n$-step empowerment, as we can determine the initial successor states in the same tree. The colours indicate code changes for the modification. For basic UCT empowerment read just the code in black. For \textit{aggregated empowerment}, add the \textcolor{red}{red} line at 18. For \textit{novelty bias} add the two \textcolor{blue}{blue} code snippets in 29 and 36. For the \textit{full branching} modification add the code in \textcolor{green}{green}, and set DEPTH to a non-zero value.}
\caption{Comparisons of our proposed approach to the state-of-the-art on PRID2011, iLIDS-VID, and MARS datasets. The rank-1 accuracies are reported and for MARS we provide mAP in brackets. The best and second best results are marked by \textcolor{red}{red} and \textcolor{blue}{blue} colors, respectively.}
\caption{Comparison of a human summary to best single- and multi-agent model summaries, (m3) and (m7) from CNN/DailyMail dataset. Although single-agent model generates a coherent summary, it is less focused and contains more unnecessary details (\colorbox{pink}{highlighed red}) and misses \textbf{keys facts} that the multi-agent model successfully captures (\textbf{bolded}).}
\caption{Referring relationships disambiguate between instances of the same category by using their relative relationships with other entities. Given the relationship \relationship{person}{kicking}{ball}, the task requires our model to correctly identify which \object{person} in the image is kicking the \object{ball} by understanding the predicate \predicate{kicking}.}
\caption{Referring relationships' inference pipeline begins by extracting image features, which are then used to generate an initial grounding of the \object{subject} and \object{object} independently. Next, these estimates are used to shift the attention using the \predicate{predicate} from the \object{subject} to where we expect the \object{object} to be. We modify the image features by focusing our attention to the shifted area when refining our new estimate of the \object{object}. Simultaneously, we learn an inverse shift from the initial \object{object} to the \object{subject}. We iteratively pass messages between the subject and object through the two predicate shift modules to finally localize the two entities.}
\caption{(a) Relative to a subject in the middle of an image, the predicate \predicate{left} will shift the attention to the right when using the relationship \relationship{subject}{left of}{object} to find the object. Inversely, when using the \object{object} to find the \object{subject}, the inverse predicate \predicate{left} will shift the attention to the left. We visualize all $70$ VRD, $6$ CLEVR and $70$ Visual Genome predicate and inverse predicate shifts in our supplementary material. (b) We also show that these shifts are intuitive when looking at the dataset that was used to learn them. For example, we find that \predicate{ride} usually corresponds to an \object{object} below the \object{subject}.}
\caption{Phase shift induced by free-streaming neutrinos and other light relics. \textit{Top:}~Template of the phase shift $f(k)$~(\textcolor{pyBlue2}{blue}) as defined in~\eqref{eq:phaseTemplate}, with the fitting function~\eqref{eq:phase_fit} shown as the \textcolor{pyRed}{red} curve. The template was obtained numerically in~\cite{Baumann:2017gkg} by sampling the phase shift in 100~different cosmologies with varying free-streaming radiation density. The blue bands indicate the $1\sigma$~and $2\sigma$~contours in these measurements. \textit{Bottom:}~Linear BAO~spectrum~$O(k)$, defined in~\eqref{eq:Pg}, as a function of the amplitude of the phase shift~$\beta$.}
\caption{Observational constraints on the amplitude of the phase shift~$\beta$. \textit{Left:}~Contours showing $1\sigma$~and $2\sigma$~exclusions in the $\alpha$-$\beta$ plane for the two redshift bins~$z_1$ and~$z_3$ from our Fourier-space analysis of the BOSS~DR12~data, both from the BAO~data alone and after imposing a CMB~prior on the BAO~frequency parameter~$\alpha$. The degeneracy between the parameters~$\alpha$ and~$\beta$ is clearly visible. By imposing a prior on~$\alpha$ from the~CMB, we restrict the values of the BAO~frequency, or equivalently the BAO~scale, to be consistent with observational constraints from the Planck satellite. \textit{Right:}~One-dimensional posterior distributions of~$\beta$ without~(\textcolor{pyBlue2}{blue}) and with~(\textcolor{pyRed2}{red}) the $\alpha$-prior from Planck for the combined redshift bins. The dashed line is the result after marginalizing over the lensing amplitude~$A_L$, which is a phenomenological parameter that exhibits a large fluctuation in the cosmology inferred from the Planck data. Even in this case, we exclude $\beta=0$ at more than 95\%~confidence.}
\caption{Validation of the modified BAO~analysis employed in this paper. The displayed posterior distributions for the amplitude of the phase shift~$\beta$ are computed in likelihood-based forecasts for scenarios in which the mock BOSS~data were generated using $N_\mathrm{eff}^\mathrm{in} = 3.046\text{ (\textcolor{pyBlue}{blue}) and }0$ (\textcolor{pyRed}{red}), corresponding to $\beta = 1\text{ and }0$. In both cases, the model in~\eqref{eq:model_Neff0} used a fiducial cosmology with $\Neff = 3.046$. The dashed lines show the posterior distributions after imposing a prior from a Planck-type CMB~experiment. We see that the posteriors reproduce the expected behavior which indicates that the estimation of~$\beta$ is essentially unbiased.}
\caption{Validation of the configuration-space analysis using mock catalogs. The left column contains a comparison of the distribution of maximum-likelihood values in the \mbox{$\alpha$-$\beta$}~plane for the redshift bin~$z_1$~(\textit{bottom}) and for $\beta$~(\textit{top}) in 999~mock catalogs~\cite{Kitaura:2015uqa} for the Fourier-space~(FS, \textcolor{pyBlue}{blue}) and configuration-space~(CS, \textcolor{pyRed}{red}) analyses. On the right, we show the correlation between the inferred phase shift amplitudes in the two analyses~(\textcolor{pyGreen}{green}).}
\caption{Observational constraints on the amplitude of the phase shift~$\beta$ from our configuration-space analysis of the BOSS~DR12~data. \textit{Left:}~Contours showing $1\sigma$~and $2\sigma$~exclusions in the plane spanned by the BAO~scale parameter~$\alpha$ and the phase shift amplitude~$\beta$ for the two redshift bins~$z_1$ and~$z_3$, both from the BAO~data alone and after imposing a CMB~prior on~$\alpha$. \textit{Right:}~One-dimensional posterior distributions of~$\beta$ without~(\textcolor{pyBlue2}{blue}) and with~(\textcolor{pyRed2}{red}) the $\alpha$-prior from the Planck satellite for the combined redshift bins resulting in $\beta_\mathrm{CS} = 0.4\pm 2.1$ and $\beta_\mathrm{CS} = 2.55 \pm 0.80$, respectively. The shift in the mean value originates from lower values of~$\alpha$ in conjunction with the discussed degeneracy between~$\alpha$ and~$\beta$.}
\caption{(a) -- a simplified scheme of the transitions in a muonic atom. (b) -- correlated P- and D-spectra measured with enriched $^{76}$Se target. Some of the identified \gray\transitions are indicated.``DE" and ``SE" denote the double and single escape peaks. The\muXray\transitions for carbon and oxygen are indicated as$\mu$X(C) and $\mu$X(O).}
\caption{Evolution of $\gamma$-lines intensity with time after $\mu$-stop in Se targets. The figure shows two separate measurements, one with an enriched $^{76}$Se target (top graphs) and one with a natural Se target. In the latter the main components of the initial Se-isotopes are indicated. The identification is based on the different lifetimes and the isotopic abundances. The deterioration of the time resolution below $\approx 200$~keV \gray\energy is clearly observed.}
\caption{The top ten ranking teams of the ISBI longitudinal MS lesion segmentation challenge (\textcolor{blue}{https://smart-stats-tools.org/lesion-challenge}) as of September 2018 with average metrics of challenge score, Dice coefficient, Jaccard coefficient, positive predictive value (PPV), sensitivity (TPR), lesion TPR based on lesion count (LTPR), lesion FPR based on lesion count (LFPR), Volume Difference (VD), Average Symmetric Surface Difference (SD) and average segmentation volume. Average manual volume of the two raters in the challenge was \textbf{15,648}. Our proposed method (IMAGINE) with $F_\beta$ achieved the best results in 4 out of 9 evaluation metrics, and second highest in DSC and Jaccard compared to the other methods, while our patch-wise 3D FC-DenseNet trained with focal loss achieved the first place as well as better results in PPV and LFPR metrics. For some of the most recent submissions, UVA, Unige, PAVIS and Braz, we could not find any published articles for reference. In case of multiple submissions by same group, we reported their highest ranked results. It is noteworthy that the performance of methods with the overall challenge score of over 90 is considered to be comparable to human raters.}
\caption{Comparison of training schedules for~\eqref{eq:losses}, evaluated on VOC2007 test. $n:\alpha:\beta$ indicates training for $n$ iterations with the given $\alpha$,$\beta$ values. See text for additional explanations. The best score in each column is colored by \textcolor{red}{red}. }
\caption{Number of classes and instances in our datasets. For PASCAL~VOC, we used all the annotations including \texttt{difficult} boxes. Note that there are twenty object classes in PASCAL~VOC and \datasetclipart, and six object classes in \datasetwatercolor~and \datasetcomic.}
\caption{ Example outputs for our DT+PA using SSD300 as the baseline FSD in the test set of \datasetwatercolor. We only showed windows whose scores are above 0.25 so as to maintain visibility.}
\caption{Relative probability of friendship and interaction 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{(Color online) Dynamical destabilization of the dark soliton stripe corresponding to the full NLS (\ref{dark1}) [see background colormap] and the AI reduction (\ref{1D_PDE_DS}) [see (green) curves in the corresponding top row in each set of panels]. % The corresponding systems are initialized with a dark soliton stripe at $x(y) = x_0 + A\cos(n \pi y/L_y)$ with $x_0=4$, $A=0.1$, with $\mu=40$, and $n=1$ (top set of panels) and $n=2$ (bottom set of panels). % Within each set of panels the top and bottom row correspond to the magnitude ($|u(x,t)|$) and phase of the field at the indicated times. We note that, for better comparison between frames, the phase has been rotated so that, for all times, the phase at the origin is fixed to $\pi/2$. % We also note that, for better visibility of the destabilization features, the panels only depict the domain for $x\geq 0$ (the $x<0$ region has trivial dynamics as there is no stripe in there). %See supplemental movies \verb|DS1_movie| and \verb|DS2_movie|. See supplemental movies {\tt DS1\symbol{95}movie} and {\tt DS2\symbol{95}movie}. }
\caption{(Color online) Dynamical destabilization of the DB soliton stripe corresponding to the full NLS (\ref{dbeom}) [see background colormap] and the AI reduction (\ref{dbeqn}) [see (green) curves in the corresponding top row in each set of panels]. % The perturbations to the initial DB soliton stripe and layout of the figure are the same as in Fig.~\ref{fig:PDE_AI_DS} with the addition of a third row of panels depicting the magnitude of the second field ($|v(x,t)|$). The values of the chemical potentials are $\mu_d=40$ and $\mu_b=29.6815$. See supplemental movies {\tt DB1\symbol{95}movie} and {\tt DB2\symbol{95}movie}. }
\caption{\textcolor{blue} Overview of our saliency learning and exploration approach.}
\caption{ \textbf{Examples of automatic ordinal labeling.} \textcolor{blue}{Blue mask}: foreground ($\Ford$) derived from semantic segmentation. \textcolor{red}{Red mask}: background ($\Bord$) derived from reconstructed depth. \label{fig:automatic_label}}
\caption{ \% {\bf MRE} and \% {\bf SA} seen in 20 repeats. {\bf Med} is the 50th percentile and {\bf IQR} is the {\em inter-quartile range}; i.e., 75th-25th percentile. Lines with a dot in the middle (e.g.,\protect\quartex{3}{13}{13}{0}) show median values with the IQR. MRE and SA results are sorted in different directions since better MRE and SA values are smaller and larger (respectively). The left-hand side columns {\bf Rank} results (and the {\em smaller}, the {\em better}). Ranks separate statistically different results, as computed by a bootstrap test (95\% confidence) and the A12 test~\cite{Whigham:2015:BMS:2776776.2738037}). \ofr denote results that are so bad, that they fall outside of this figure s range of [0,100] \%. \colorbox{black!10}{1*} denotes rows of faster best-ranked methods.}
\caption{Recall scores on the held out set of 1,000 images/captions for the three matchmap similarity functions. We also show results for the baseline models which use automatic speech recognition-derived text captions. \textcolor{red}{TODO: clean up the table here, verify that the text follows the table well}}
\caption{ROC curves for scores obtained when matching intra-session samples within the first acquisition session (\textcolor{myblue}{\bf NIR images: blue line}, \textcolor{myred}{\bf R images: red line}, \textbf{cross-wavelength matching: dashed black line}). Equal error rate (EER) is also shown for each case. Plots generated only for successful comparisons (\ie not rejected by the software due to low image quality).}
\caption{Long-term analysis for \textbf{all subjects} for the \textbf{OSIRIS} matcher using automatic iris segmentation, for both NIR and R samples. \textcolor{mygreen}{\bf Green dots represent correct behavior, \ie a match for genuine pair, and a non-match for impostor pair}, while \textcolor{myred}{\bf red dots correspond to incorrect behavior, \ie a false non-match for a genuine pair, and a false match for impostor pair} between samples acquired in the first session (after 5-7 hours) and samples acquired in the following sessions.}
\caption{log10(1/$\mu_{1}$) for the network shown in Fig.~\ref{fig:network1} as we increased the number of nodes subject to attacks. (\protect\markerone) symbols: Defenders are placed on nodes 1, 3 \& 10 ; (\protect\markerfour) symbols: Defenders are placed on nodes 2, 7 and 9. Attackers are chosen in a random order, but ensuring that no node is both an attacker and a defender.}
\caption{(a) Plot of $\bm{n}^T\bm{w}_1$ vs the degree of the attacked node for a 300 scale free network with average degree 2. (b) Plot of $\beta^2$ vs the degree of the attacked node. (\protect\markerthree) symbol indicates attacked nodes with $\Delta=3$. (\protect\markerone) symbol indicates attacked nodes with $\Delta=2$. (\protect\markertwo) symbol indicates attacked nodes with $\Delta=1$. We perform calculations setting the final time $t_{f}$=1 with $s_i = 2.5$, $r_i = 0$ and $\epsilon =10^{-2}$. }
\caption[]{3D pole, crossarm and wire (\textcolor{black}{$\blacksquare$}) geometry and parameters including example protection zones (\textcolor{red}{$\square$}) and arbitrarily located platform. }
\caption[]{2D ($xy$-planar view) pole, crossarm and wire parameters including example protection zones (\textcolor{red}{$\square$}) and crossarm region (\textcolor{black}{$--$}). }
\caption[]{Example crossarm region coverage for $\mathbf{p}_q(2,0,4)$, $\phi=\theta=\psi = 0$ and $\mathbf{u}=\mathbf{0}$. The RealSense field of view (\textcolor{gray}{$\square$}) using $H_C,V_C,R_C$ and crossarm region $\partial \mathcal{C}$ (\textcolor{red}{$\square$}) are shown along with the sensor coverage of the $xy$-plane at pole altitude $\partial \mathcal{S}_{xy}$ (\textcolor{blue}{$\square$}), sensor coverage of the crossarm region $\partial \mathcal{S}_{c}$ (\textcolor{green}{$\square$}) and sensor coverage of the pole region $\partial \mathcal{S}_{d}$ (\textcolor{green}{$\blacksquare$}).}
\caption[]{Example crossarm region coverage for $\mathbf{p}_q(0.5,0,z)$, $\phi=\theta=\psi = 0$, $\mathbf{u}=\mathbf{0}$ with $z\in\{-2.5,-2.0,-1.5,-1.0\}$ (clockwise from upper left) and RealSense field of view $H_C,V_C,R_C$. The crossarm region $\partial \mathcal{C}$ (\textcolor{red}{$\square$}) is shown along with the sensor coverage of the crossarm region $\partial \mathcal{S}_{c}$ (\textcolor{green}{$\square$}) and sensor coverage of the pole region $\partial \mathcal{S}_{p}$ (\textcolor{green}{$\blacksquare$}).}
\caption[]{Detection area metric $A_s/A_p$ with $\mathbf{p}_c(0, 0, 6)$, $r_c = 1.5$, $r_d=0.5$ for $r_q \in (0.5,5.0)$ and $\Delta h \in (-4,4)$. As the detection area metric goes from zero (\textcolor{red}{$\blacksquare$}) to unity (\textcolor{blue}{$\blacksquare$}), the sensor covers more of the pole region such that the probability of detected the crossarm increases, regardless of the crossarm orientation $\psi_c$.}
\caption[]{Flight mode transition architecture where one and two-way mode transitions are shown as well as the inspect sub-modes (\textcolor{blue}{$\blacksquare$}).}
\caption{SSE for various temperature gradients of a) \eps-70~nm, b) \epsal-70~nm and c) \eps-40~nm. Insets show the dependence of SSE at 0~kOe ($\triangle$) and 40~kOe (\textcolor{red}{$\nabla$}). d) Schema of the longitudinal experimental configurations. Directions of spin current ($\vec{J}_s$), temperature gradient ($\nabla\vec{T}$), magnetization ($\vec{M}$), and electrical field resulted from inverse spin Hall effect ($\vec{E}_{ISHE}$), are shown. }
\caption{The design process of the material at different iteration steps. \break \protect \newboxsymbol{magenta}{magenta} Young modulus of $1.82$, \protect \newboxsymbol{cyan}{cyan} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{The design process of the material at different iteration steps. \break \protect \newboxsymbol{magenta}{magenta} Young modulus of $1.82$, \protect \newboxsymbol{cyan}{cyan} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{The design process of the material at different iteration steps. \break \protect \newboxsymbol{magenta}{magenta} Young modulus of $1.82$, \protect \newboxsymbol{cyan}{cyan} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{The design process of the material at different iteration steps. \break \protect \newboxsymbol{magenta}{magenta} Young modulus of $1.82$, \protect \newboxsymbol{cyan}{cyan} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{Thermocouple fabrication stages and patterns of electrodes at different magnifications. Thermocouples were fabricated on glass and Si$_3$N$_4$ membranes of different thicknesses: 1~$\mu$m and 30~nm. (a) Metal junctions of 2.5-$\mu$m-wide metal stripes with $100\times 100~\mu$m$^2$ primary contact pads. (b) Photo image of a laser ablated photolithography mask \blue{used} for resist exposure. It defines the secondary contact pads to interface with electrical measurements. (c) Photo of the final device on glass. (d) \blue{A} SEM image of the micro-thermocouple \blue{and reference electrodes}. The central pair is \blue{the} Au-Ni thermocouple.}
\caption{Characterisation of thermocouples. (a) Temperature increase induced by laser heating at different laser power\blue{s} measured by the \blue{optical modulation} method (wavelength $\lambda = 830$~nm; $p$-polarisation at slanted front-side incidence). Thermocouple Au-Ni was made on a slide glass. \blue{Sensitivity of} $10.1~\mu$V/K~\cite{Chu1} \blue{determined for similar thermocouple was used for estimation of temperature changes}; Au-Au junction was used as a reference. Illumination of the substrate was \blue{carried out} from the side to contacts. (b) Temperature vs. laser power for 1~$\mu$m-thick Si$_3$N$_4$ membrane. \blue{The Au-Au reference electrodes had a Cr adhesion layer and formed a thermocouple which was experiencing a thermal gradient due to asymmetry of the primary contact pads during laser heating (see panel (c)).} (c) An optical see-through image of the $400\times 400~\mu$m$^2$ SiN-membrane region with thermocouple whose response is plotted in (b); laser spot was $\sim 100~\mu$m in diameter. \blue{Note a thermal asymmetry of this layout where the upper $100\times 100$~$\mu$m$^2$ primary contact square pad was on the SiN membrane while the lower one on the Si substrate.}}
\caption{Thermocouple on a 30-nm-thick SiN-membrane. (a,b) SEM images of thermocouple made \blue{on a} $250\times 250~\mu$m$^2$ SiN window. Note, secondary (large) contact pads are made from the same metal (Au or Ni) as the smaller ones to avoid formation of a secondary thermocouple. The large contacts are placed on Si substrate to \blue{remove a} thermal gradient on the thermocouple \blue{(such gradient was responsible for the observed temperature change in Fig.~\ref{f-thick}(b) measured with the Au-Au contact)}. (c) Temperature vs. laser power. (d) Temperature vs. \blue{position of the} electron beam across the central cross section (along the line in (b)). Diameter of the e-beam \blue{was} $\sim 0.5~\mu$m at the acceleration voltage of 25~keV; modulation frequency was 30~Hz, current $\sim 1.7$~nA as measured by Faraday cup without \blue{the} sample. Central shaded region depicts \blue{the} location of membrane. E-beam was scanned across the SiN window and \blue{Si} substrate without \blue{direct} exposure of metal leads/contacts.}
\caption{Temporal response of thermocouple on a 30-nm-thick SiN-membrane to a square-wave optical excitation. (a) Video image of a tightly-focused laser beam onto thermocouple with $\sim 10~\mu$m spot diameter; $\lambda = 830$~nm. (b) Temporal response of thermocouples: (\emph{i}) to a 3.6~mW laser power at repetition rate $f = 2$~kHz with a thermocouple on a slide glass and (\emph{ii}) to 1.8~mW power with thermocouple on \blue{a} 30-nm-thick SiN-membrane at $f = 2, 5, 10$~kHz; note different x-axis scales in (b). The fastest switching time was $\tau= RC = 10~\mu$s with ohmic resistance of thermocouple $R = 500~\Omega$ and $C = 20$~nF. Electronic pre-amplifier of 100$^\times$ was used for a direct observation by oscilloscope. The \blue{estimated} max-min $\Delta T$ span was 16~K (30.7-to-14.7~K above RT of 22$^\circ$C) for the thermocouple on glass and $\Delta T = 4.1$~K for 30~nm SiN-membrane (31.3-to-27.2~K) at 2~kHz; at higher 5 and 10~kHz frequencies the max temperature increase was similar $\Delta T_{max} = 30.5$~K and min-max span of $\sim$4~K. }
\caption{Response of micro-thermocouple to back-side electron irradiation. (a) SEM image of thermocouple on 30~nm SiN-membrane by back-scattered (in lens) and secondary (large angle \blue{scattered}) electrons. (b,c) Measured amplitude and phase response of the Au-Ni thermocouple to a diagonal scan (dashed line in \blue{the} inset in (b)) with $\sim 7~\mu$m steps measured with a lock-in amplifier. Thermocouple voltage was normalised to the transmitted electron current measured by the Faraday cup using an additional lock-in amplifier; e-beam blanking frequency was 27~Hz. The slope of line \emph{(1)} in (c) corresponds to the best fit by a linear $Phase\sim Const\times d$ dependence, where $d$ is the distance between heat source and measurement point (only valid at large separation\blue{s}); the temperature diffusivity was $a = 0.62\times 10^{-4}$~m$^2$/s or 47.7 times larger than that of SiN~\cite{Grigoropoulos}.}
\caption{ A particle in the reference camera (\textcolor{blue}{circle}) can lead to multiple epipolar-consistent putative matches (\textcolor{red}{circles}). However, only a subset of them represents true 3D particles (\textcolor{green}{triangles}). \emph{Left}: 1D-illustration. \emph{Right}: peak in reference camera (0.1ppp). \emph{Bottom:} other camera view with 5 putative matches that are consistent over all 4 cameras. Moreover, one true particle has initially not been found, due to visually overlapping ones that bias peak detection; but will be found in subsequent iterations, when the overlapping particles have been reconstructed and removed from the residual images or the triangulation threshold has been relaxed. }
\caption{Speed of each additional walker. Green stars represent \eqref{Eq: Multiple}, red markers are experimental data points with error bars\cite{FHV15}, and blue triangles are from the algebraic model of Filoux \ea \cite{FHV15} with the dashed curve fitting the points from their model. For the green stars $\gamma = 0.44$ and $\kappa = 1/7$ from \eqref{Eq: Multiple}. Further, the distance between droplets, $s = \theta - \theta^1 = (8/6)\cdot 2\cdot \pi/\omega \approx 8 mm$, similar to the spacing used in \cite{FHV15}. The markers from the results of this article were embedded directly onto the original figure by precisely matching the axes, and without further modification or manipulations. {\color{red} Figure adapted and modified with permission from Filoux \ea \cite{FHV15}}.}
\caption{Ratio of speeds of a pair of droplets as the separation is increased. Green stars represent \eqref{Eq: Multiple}, red markers are experimental data points with error bars\cite{FHV15}, and blue triangles are from the algebraic model of Filoux \ea \cite{FHV15} with the dashed curve fitting the points from their model. For the green stars $\gamma = 0.44$ and $\kappa = 1/7$ from \eqref{Eq: Multiple}. \textbf{(a)} Ratio of speeds without offsets or rescaling of axes as presented in the original figure of Filoux \ea \cite{FHV15Arxiv}. \textbf{(b)} Ratio of speeds offset by unity with a log-scale for the ordinate as presented in the original figure of Filoux \ea \cite{FHV15}. The markers from the results of this article were embedded directly onto the original figures by precisely matching the axes, and without further modification or manipulations. {\color{red} Figures adapted and modified with permission from Filoux \ea \cite{FHV15, FHV15Arxiv}}.}
\caption{Data2Vis qualitative evaluation interface with results from beam search. (a) A user can load a random dataset from the RDdataset collection or paste a dataset (JSON format) and select ``Generate.'' (b) User can paste a JSON dataset and select "Generate"(c) Data2Vis generates Vega-Lite specifications using beam search (beam width = 15 in this case) based on the dataset. The user can modify and iterate on any of the visualizations using the Vega-Lite editor. Highlights below each visualization represents cases of valid specifications \protect\greenbox\xspace and incomplete specifications \protect\orangebox\xspace where the model attempts to use variables not in the dataset (phantom variables). \label{fig:vizinterface}}
\caption[]{\textbf{Sketch for learning a pose specific representation from unlabeled data.} We learn to predict a low-dimensional latent representation and, subsequently, a different view of the input, \emph{solely} from the \textcolor{blue_ours}{latent representation}. The error of the view prediction is used as \textcolor{green_ours}{feedback}, enforcing the latent representation to capture pose specific information without requiring labeled data. }
\caption[]{\textbf{Architecture sketch for semi-supervised learning.} The input view, $\mathbf{x}^{(1)}$, is mapped to the \textcolor{blue_ours}{latent representation}, $\theta$, by the encoder $f_{1}$. Based on $\theta$, the decoder $g_{2}$ is required to generate a different view $2$ of the input. At the same time the latent representation is ensured to suffice a linear mapping, $g_{l}$, to the 3D joint positions by employing labeled samples. This is illustrated by the green paths depicting the \textcolor{green_ours}{gradient flow} to the latent representation and, consequently, to the encoder.}
\caption{\blue{{Comparison of the main figures of merit of TFETs made of several 2D channel materials. NW, DG, SG and GAA stand for nanowire, double-gate, single-gate and gate-all-around, respectively.}}}
\caption{\blue{\textbf{Local intrinsic Fermi level, charge, and transmission coefficient of the stanene nanoribbon.} 2D colormap of a) the free charge distribution and b) the intrinsic Fermi energy, c) transmission coefficient at the intrinsic Fermi energy as a function of the position of the nanoribbon, for three different bias conditions: $p$-conductive ON state (top), OFF state (center) and $n$-conductive ON state. V$_{\text{OFF}} = 0$ V, i.e. the minimum of the transfer characteristic in d). A sketch of the TFET band diagram is depicted aside the intrinsic Fermi level colormap. The actual width of the nanoribbon is marked by dashed lines. A lateral oxide of $1$ nm is considered at both sides of the nanoribbon. d) Transfer characteristic of the TFET in linear and semi-logarithmic scales.} }
\caption{\blue{\textbf{Impact of channel-length scaling on the 1D-channels TFET.} Transfer characteristics in a) linear and b) logarithmic scale of the TFET for different gate underlaps: $22.6$ nm (black diamonds), $16.9$ nm (red squares), $11.3$ nm (blue circles), and $5.6$ nm (gray triangles).}}
\caption{Peak intensity maps for HCN \JJ{3}{2} (left) and \methanol $11_0$--$10_1$ A$^{++}$ (right), calculated from data smoothed into $2\ \kmps$ and $36\ \kmps$ velocity bins, respectively. The HCN \JJ{3}{2}\intensities are measured for the velocity range$-100$ to $-70$\\kmps. The shape of the $-80$-\kmps\filament is indicated in the HCN map, with two prominent features, labeled knot-A and knot-B. The crosses on the\methanol\map are the positions of the averaged spectra used in Figure\ref{FIG_SPECALL} and the population diagram analysis (\S\ref{ROT_ANALY}). An approximate shape for the hypothesized gas stream (read from Figure 3 of \cite{Oka2016}) is indicated by the dashed lines.\label{FIG_FILAMENT}}
\caption{Maps of (a) moment-0, (b) moment-1, (c) velocity widths (\Dv), and (d) moment-1 smoothed with $20''$ Gaussian for the cloud image reconstructed from the six main branches and their common trunk nodes. The width \Dv\is defined as the moment-1 value divided by the spectral peak intensity toward each line of sight. The 2D spatial distributions of the branches are shown by the contours overlaid on the moment-0 map. The shape of the$-80$-\kmps\is denoted by the dashed-lined curves on the moment-1 and\Dv\maps, which are identical to that in Figure\ref{FIG_FILAMENT}. The rectangles in the moment-1 map indicate the strips for the PV diagrams in Figures \ref{FIG_STRIPS} and \ref{FIG_ERROR}. The approximate shape of the hypothesized gas stream \citep{Oka2016} is shown in panel (d).\label{FIG_DGRAM3}}
\caption{(a) Rotational population diagram for \methanol\and\methanimine\in the$-80$-\kmps\filament. (b) Physical conditions in the filament calculated from a non-LTE analysis using the methanol lines. Credible intervals obtained from a flat-prior Bayesian analysis are shown by the colored contours.\label{FIG_ROT}}
\caption{(a) \COt\\JJ{3}{2}\integrated intensity map for the$-110$ to $0$\\kmps\velocity range obtained with the ASTE 10-m telescope\citep{Tanaka2014}, with contours of the ALMA \JJ{3}{2}\integrated flux density drawn at 5, 10, 20, 50, and 100 Jy\,$\mathrm{beam^{-1}}\cdot\kmps$. (b) PV diagrams for the ALMA HCN \JJ{3}{2}\and ASTE\COt\\JJ{3}{2}\data (left and middle panels, respectively) for the rectangular field in panel (a), where the$X$ axis is in the direction of the long side of the rectangle, and the spectra are averaged in the direction of the short side. \textcolor{red}{The bright emission band in the 15--$20\ \kmps$ velocity channels of the \COt\diagram is the foreground infrared dark cloud G359.62{$-0.24$}.} The HCN and \COt\images are overlaid in the rightmost panel.\label{FIG_SINGLEDISH_13CO}}
\caption{Dendrogram for the clumps identified in the HCN \JJ{3}{2}\PPV map. Leaf clumps growing directly from the ground level (``sprouts'') are omitted. The thick-lined trees represent the\theObj\main body. The six main branches are colored other than black.\label{FIG_DGRAMTREE}}
\caption{\textbf{Trajectory estimates} on a sequence from the 7-Scenes test split. Top (left to right): Output trajectories are shown in \textcolor{red}{red}, against ground-truth trajectories in \textcolor{blue}{blue}. Bottom: $\se3$ estimates of relative poses. Each of the $6$ $\se3$ coordinates is plotted independently. On this sequence, CTCNet performs better than LSTM$_{gt}$.}
\caption{\textbf{Outlier detection}: Upon covariance recovery, estimates with covariance above a threshold are marked outliers (here shown in \textcolor{blue}{blue}).}
\caption{NEP performance (HR@10) in cold-start regime as a function of training frequency of the pair (query item, next item) on 30Music and Deezer datasets. \textcolor{red}{}}
\caption{Different regimes of gelation. (a) Phase diagram. Experimental points are categorised based on the final state obtained in the reservoir cell. The spinodal line is obtained from free volume theory in polymer dilute regime. The dashed line is the polymer overlap concentration in the free volume. State points analysed in b and c are circled. (b) Growth of the characteristic wave number. By increasing density: $\phi=4.2,\,8,\,16,\,27~\%$, $c_p=1,\,1.5,\,1.2,\, 1$ mg/g for \textcolor{red!40!yellow}{$\blacklozenge$}, \textcolor{red!80!yellow}{$\circ$}, \textcolor{red!80!black}{\tiny$\blacksquare$} and \textcolor{red!40!black}{$\bullet$} respectively. The lines are possible scaling laws for the intermediate coarsening regime. (c) Comparison of system evolution in terms of largest cluster extent and of mean coordination number for the same samples.}
\caption{Segmentation results on person category of Coco val2017. AP\textsubscript{S}, AP\textsubscript{M}, and mAP\textsubscript{L} is the mean Average Precision for small (area $< 32^2$), medium ($96^2 >$ area $> 32^2$) and large (area $> 96^2$) objects respectively. Note that our result is tested on Coco val2017 while other results on the Coco detection leaderboard are tested on test2017. \color{red}The results would be update after iterations of training.}
\caption{Model overview. Our image-to-image translation model consists of two auto-encoders~(denoted by \textcolor{myred}{red} and \textcolor{myblue}{blue} arrows respectively), one for each domain. The latent code of each auto-encoder is composed of a content code $c$ and a style code $s$. We train the model with adversarial objectives~(dotted lines) that ensure the translated images to be indistinguishable from real images in the target domain, as well as bidirectional reconstruction objectives~(dashed lines) that reconstruct both images and latent codes. % See Sec.~\ref{sec:model} for more details. }
\caption{{\color{black!50!black}The energy measurement error of EMPIOT versus the ground truth. Error bars show median, lower quartile and higher quartile. Each marker is the median of ten experiments each 30 seconds long.} }
\caption{Benchmark results of forward convolution of \alexnet's ``conv2'' layer on \phundred. We use 64 MiB workspace size and a mini-batch size of 256. Numbers on each rectangle represent micro-batch sizes.}
\caption{Benchmark results of \alexnet \on three different GPUs with different workspace sizes (8, 64, 512 MiB). The labels ``u'', ``p'' and ``a'' represent\ttundivided, \ttpoweroftwo, and \ttall, respectively. We use a mini-batch size of 256 on \keighty \and\phundred, and 1024 on \vhundred.}
\caption{Benchmark results of different CNNs on \phundred \with different workspace sizes (8, 64, 512 MiB), using\tensorflow \framework. We use a mini-batch size of 256 for\alexnet \and\densenet, and 64 for \resnet-50.}
\caption{ Per-layer breakdowns of memory consumption of \alexnet \and\resnet-18 on \phundred. For simplicity, we only show the memory usage of unique convolutional layers (CONV\_$n$) and fully-connected layers (fc or fc$n$) in one forward propagation. We use a mini-batch of 256 for \alexnet \and 128 for\resnet-18 respectively. We set a per-layer workspace limit of 512 MiB for \cudnn, and 64 MiB for \ucudnn. Each bar segment of ``WS (\ucudnn)'' represents the maximum workspace size of the layer. }
\caption{Benchmark results of \alexnet \and\resnet-50 on \phundred \with different workspace sizes and policies (\wreuse \and\wdiv). We use a mini-batch size of 256 for \alexnet \and 32 for\resnet-50. Note that the adjoined bars have the same workspace limit in total.}
\caption{Assigned workspace division of \alexnet \on\phundred. ``F'', ``BF'', ``BD'' represent kernel types (Forward, BackwardFilter, BackwardData respectively). We use a mini-batch size of 256 for \alexnet. We set a workspace limit of 8 MiB for \wreuse, and a total workspace limit of 120 MiB for \wdiv.}
\caption{Desirable configurations (i.e. a Pareto front) of \alexnet's ``conv2'' layer (Forward) on \phundred \with a maximum workspace size of 120 MiB, and a mini-batch size of 256. Colored bars corresponding to data points represent the division of the mini-batch and the chosen micro-batch algorithms. For example, the top-left point divides the mini-batch into two micro-batches of 128 and utilizes the{\tt FFT\_TILING} algorithm.}
\caption{Architecture of our RecNet. It takes $I^w$ and $I^d$ as input to produce the restoration result $\hat{I}$. Reconstruction loss and global adversarial loss are adopted across entire image (labeled in \textcolor[RGB]{112,48,160}{\textbf{purple}}), while local adversarial loss is adopted across face region (labeled in \textcolor[RGB]{0,176,80}{\textbf{green}}).}
\caption{Quantitative results on two test subsets. Numbers in the parentheses indicate SSIM and the remaining represents PSNR (dB). The best results are highlighted in {\bf \color{red}red} and second best ones except our GFRNet variants are highlighted in {\color{blue}blue}. }
\caption{Quantitative evaluation of state-of-the-arts CS reconstruction algorithms: Average PSNR$\backslash$SSIM$\backslash$time$\backslash$network Layers for sampling ratios 0.01, 0.02, 0.1 on dataset Set5. {\color{red}Red} text indicates the best and {\color{blue}blue} the second best performance}
\caption{Quantitative evaluation of state-of-the-arts CS reconstruction algorithms: Average PSNR$\backslash$SSIM$\backslash$time$\backslash$ network Layers for sampling ratios 0.01, 0.02, 0.1 on dataset Set14. {\color{red}Red} text indicates the best and {\color{blue}blue} the second best performance}
\caption{Benchmark results. %We omitted the examples which do no require any mover checks i.e., all the write instructions have no reachable buffer-free read instructions before a fence. \textcolor{red}{The second column (SV) states whether the original program (without read abstractions) reach to the same shared variable states. SR column shows whether the original program is robust according to the standard definition or not.} The fourth column (RB) stands for the robustness status of the original program according to our extended $hb$ definition. RA column shows the number of read abstractions performed. RM column represents the number of read instructions that are checked to be right movers and the LM column represents the write instructions that are shown to be left movers. PO shows the total number of proof obligations generated and VT stands for the total verification time in seconds.}
\caption{(A) Topological phase diagram for a triangular wire with $V_{\rm eff}(\ell)=0$ and $\phi_\ell=0$. The white areas are topologically trivial and the orange regions are nontrivial. The 4-star symbols indicate gapless superconducting phases. (B) Topological phase diagram for a triangular wire with $V_{\rm eff}(\ell)\neq 0$ and $\phi_\ell=0$. The values of the effective potential on the 6 chains are $ (0.67, 0.17, -0.33, -0.33, -0.33, 0.17)$ meV. The evolution of the (minimum) quasiparticle gap along the cuts I (blue lines) corresponding to $\mu=-5.4~$meV and II (red lines) corresponding to $\mu=-4.4~$meV are shown in Fig. \ref{Fig2TT} and Fig. \ref{Fig3TT}, respectively. \blue{See also Ref. \cite{Manolescu17}} }
\caption{Dependence of the minimum quasiparticle gap on the Zeeman field along the blue cuts (I) corresponding to $\mu=-5.4~$meV in Fig. \ref{Fig1TT}. {\em Top}: $V_{\rm eff}(\ell)=0$, see Fig. \ref{Fig1TT}(A). {\em Bottom}: $V_{\rm eff}(\ell)\neq 0$, see Fig. \ref{Fig1TT}(B). The white/orange regions correspond to the trivial/nontrivial phases shown in Fig. \ref{Fig1TT}. Note the gapless superconducting phase marked be the 4-star symbol (top panel). \blue{See also Ref. \cite{Manolescu17}}}
\caption{Dependence of the minimum quasiparticle gap on the Zeeman field along the dark red cuts (II) corresponding to $\mu=-4.4~$meV in Fig. \ref{Fig1TT}. {\em Top}: $V_{\rm eff}(\ell)=0$, see Fig. \ref{Fig1TT}(A). {\em Bottom}: $V_{\rm eff}(\ell)\neq 0$, see Fig. \ref{Fig1TT}(B). The white/orange regions correspond to the trivial/nontrivial phases shown in Fig. \ref{Fig1TT}. Note the gapless superconducting phase marked be the 4-star symbol (top panel). \blue{See also Ref. \cite{Manolescu17}}}
\caption{(A) \blue{Position dependence of the normalized disorder potential along the edge $\ell=3$ of a triangular wire for a specific disorder realization. The disorder profiles along the edges $\ell=1,5$ (not shown) are different, but characterized by similar qualitative features. In particular, the characteristic length scale for the potential variations is $\delta_d=60~$nm. (B) Dependence of the low-energy spectrum on the amplitude $V_{\rm max}$ of the disorder potential for the disorder realization shown in panel (A). (C) Low-energy spectrum averaged over $50$ different disorder realizations as a function of $V_{\rm max}$. The parameters of the system are: %the same as in Fig. \ref{Fig8TT} (lower panel): wire length $L = 2.25~\mu$m, chemical potential $\mu=-5.4~$meV, effective potential $V_{\rm eff} =(0.67, 0.17, -0.33, -0.33, -0.33, 0.17)~$meV, superconducting phases $\phi_1=0$, $\phi_3=\pi/2$, $\phi_5=-\pi/2$ and Zeeman field $\Gamma_B=0.35~$meV.} }
\caption{Phase \blue{boundaries} for the triangular wire in the corner-state domain. The color code describes the minimum gap of the BdG spectra for all wave vectors. \blue{The character of each phase can be identified by counting the boundary crossings along a vertical line, starting at zero magnetic field, i.e., topological or trivial for an odd or an even number of crossings, respectively.} Along these boundaries the gap closes at $k=0$. Starting from any point outside the (A) All superconductor phases are equal to zero. (B) Phases are: 0 at one corner and $\pm \pi/6$ at the other corners, i.e. $\theta=\pi/6$ in Fig.\\ref{Sample}(A). (C) The same phase distribution, with $\theta=\pi/2$. }
\caption{Phase \blue{boundaries} for the square wire in the corner-state domain. The color code describes the minimum gap of the BdG spectra for all wave vectors. \blue{The topological or trivial character of the phases can be identified by the number of boundary crossings, as described in the caption of Fig.\\ref{Pdtri}.} (A) The superconductor phases equal to zero. (B) The superconductor phases are zero and $\theta=\pi/2$, and the electric field perpendicular to the superconductors, see Fig.\\ref{Sample}(B). (C) Again $\theta=\pi/2$, but with the electric field parallel to the superconductors. }
\caption{Sample selection quality on FCVID training set. (a) precision/recall curve. (b) transition of \textcolor[rgb]{0.23,0.49,0.71}{recall (blue)} and \textcolor[rgb]{0.99,0.55,0.14}{precision (orange)}. (c) transition of mAP on training set}
\caption{Selected samples of `cello performance' class in FCVID. (a) Regions of selected samples are shaded: \textcolor[rgb]{0.22,0.54,0.85}{max (blue)}, \textcolor[rgb]{0.22,0.66,0.22}{average (green)}, and \textcolor[rgb]{0.86,0.73,0.26}{product (yellow)}. (b) Examples that are hard for only one modality. The upper and lower row shows examples with high confidence for metadata and concat modality, respectively (upper left and lower right samples in (a), respectively). }
\caption{Top retrievals on FCVID dataset. Retrievals were performed by using the classifier trained with the webly label, where the results are sorted by classification score and marked as \textcolor[rgb]{0.8, 0.2, 0.2}{negative (red)} or \textcolor[rgb]{0.2, 0.5, 0.2}{positive (green)}. }
\caption{Example of mode asymmetry in the solar power spectrum for a mode of degree $l=0$. %, after removing the influence of the \mbn{adjacent} non-radial modes. Gray is the original spectrum. Black is \red{the} box-car smooth\red{ed spectrum, with a kernel} of width equal to the mode width. Blue is the best fit. The Sun has a positive asymmetry $\chi$.}
\caption{\textbf{Top.} Asymmetry coefficient $B_{n,l=0}$ and $B_{n,l=1}$ from TT98 %\cite{Toutain1998} for the three Virgo/SPM \red{channels} (Blue, Green, Red). The best fit (solid line) from our analysis with a constant %$\chi \propto B_{n,l}$ $\chi \propto B_{n,l}/\Gamma_{n,l}$. The $68 \%$ confidence interval is also shown (dashed lines). \textbf{Bottom.} Height-to-Noise for $l=0$ modes.}
\caption{\red{\textbf{Left.} Asymmetry coefficient $\chi$ as a function of the $\log(g)$. Colors indicate the importance of the asymmetry for {\it Kepler} stars (circle). The asymmetry measured for the Sun is the filled triangle (photometry) and the filled square (velocity). \textbf{Right.} Asymmetry coefficient as a function of the effective temperature $T_{\mathrm{eff}}$.}}
\caption{\textbf{Top.} Asymmetry coefficient $\chi$ as a function of the large separation $\Delta\nu$. Below $\Delta\nu \approx 120$ $\mu$Hz, the asymmetry is mostly negative. For higher $\Delta\nu$ the asymmetry is positive like for the Sun in photometry (triangle). \red{The square shows the asymmetry coefficient of the Sun observed in velocity.} \textbf{Bottom.} Mean frequency difference $\overline{\delta\nu}$ %for the central mode frequency between models with and without asymmetry.}
\caption[Quantitative and Qualitative Analysis of cellSTORM]{\textbf{Quantitative and Qualitative Analysis of cellSTORM} Results after summing all frames processed by the NN a) and directly coming from the camera b). When processing the video-sequence in ThunderSTORM, it introduces a checkerboard-like pattern shown in the twofold zoomed version of the yellow box in e), which can be reduced by adjusting the peak intensity threshold (e.g. $3\cdot std(frame)$) illustrated in d). c) shows the NN's result successfully compensating for the pattern effect, due to high noise and compression of the video stream. Scalebar = $5\,\mu m$. The graphs on the right hand side visualize a comparison of the achieved localization accuracies of our NN and ThunderSTORM applied to simulated data. We varied the number of photons per emitter (1000, 500, 100, 50) as well as the compression ratio of the H.264 codec (70-100\%), before the video-stream was localized by the NN and ThunderSTORM. We estimated the accuracy by measuring the Euclidean distance between a nearest neighbor in a GT and reconstructed frame and calculate the mean over all distances, visualized as green (NN) and blue (ThunderSTORM) plots. The green (NN) and blue (ThunderSTORM) bars indicate the number of correctly detected emitters within the allowed range of 200\,nm compared to the 65.489 GT events (orange). It can be seen, that the NN always detects more good quality emitters, but with a loss of accuracy at lower intensities (i.e.$\leq500\;\frac{photons}{emitter}$) compared to ThunderSTORM. }
\caption[Mean-variance plot HUAWEI P9]{\textbf{Mean-variance plot HUAWEI P9} Mean-variance plot generated using a series (ISO=3200) of unprocessed raw images (”snap-mode”) acquired by a Huawei\textregistered\P9 camera (blue points). The camera gain is constant up to an critical intensity of 220 ADU, which should not be exceeded in the experiment.}
\caption[Mean of a dark image over time]{\textbf{Mean of a dark image over time} Mean of a dark image acquired with a Huawei\textregistered\,\,P9 camera and compressed with a H.264 encoder as a function over time. The overall decay of the signal and the periodic drop of the signal demonstrates the disadvantages of a compressed video signal. Equidistant gray lines (vertical) with 1.07\,s (i.e. each 31st frame) at$29\,fps$ spacing show the periodicity of the drop in the intensity signal. Certain pixel values less than a threshold are clipped to zero.}
\caption{ Comparison of several spectral observables between the $\mathrm{SI}$ and the $\mathrm{SR}_\Omega$ models in both situations of energy cascade [$\nu=10^{-6}$ (\textcolor{PlotOrange}{\large{$\bullet$}}) and $\langle \alpha_2 \rangle \simeq 10^{6}$ (\textcolor{PlotOrange}{\large{$\circ$}})] and quasiequilibrium [$\nu=10^{-12}$ (\textcolor{PlotBlue}{\small{$\blacktriangle$}}) and $\langle \alpha_2 \rangle \simeq 10^{12}$ (\textcolor{PlotBlue}{\scriptsize{$\triangle$}})]: (a) energy spectra, (b) energy flux (\ref{eq:sabra_standard_flux}), (c) skewness $\mathcal{S}_n$ (\ref{eq:skewness}) (the inset shows the same plot with the logarithmic $y$ axis), and (d) flatness $\mathcal{F}_n$ (\ref{eq:flatness}). Errors are the order of or smaller than the symbol size. The dashed line labeled $k_n^{-0.72}$ in (a) represents the scaling behavior in the manner of Kolmogorov plus intermittency correction. The dashed line labeled $k_n^{-1}$ in the inset of (c) represents a dimensional prediction valid at quasiequilibrium; indeed, since $\langle|u_n|\rangle$ and $\langle \Pi_n^E \rangle$ do not depend on the wave number $k_n$, at least in a certain range of scales, as shown in (a) and (b) respectively, one has that $\mathcal{S}_n \sim k_n^{-1}$ [Eq.~(\ref{eq:skewness})]. The dashed line labeled $k_n^{0.06}$ in (d) shows a best fit of the curves in the cascade regime. Finally, the horizontal dashed line in (d) displays the value $\mathcal{F}_n = 2$, which is expected for complex Gaussian variables. In these figures, and in some of the following ones, to ease the identification of the various curves and avoid the superposition of different symbols, not all data points have been marked by a symbol. }
\caption{ Probability density function of the time-dependent viscosity $\alpha_2$ for the $\mathrm{SR}_{\Omega}$ model in three different cases: a situation of energy cascade [$N=20$ and $\Omega\sim10^{4}$ (\textcolor{PlotOrange}{\large{$\bullet$}})], a situation of quasiequilibrium [$N=10$ and $\Omega\sim10^{8}$ (\textcolor{PlotBlue}{\scriptsize{$\blacksquare$}})], and a case in between [$N=15$ and $\Omega\sim10^{4}$ ($\blacktriangledown$)]. The insets on the left show the corresponding typical time evolutions of $\alpha_2$. The insets on the right show the corresponding energy spectra. }
\caption{Two-dimensional use spaces \citep{Tuggy1993,Zlatev03p447} in two time periods with a target word $w$ undergoing innovative meaning change. Dots represent uses of $w$. {\color{gray} Spatial proximity} of two uses means high relatedness.}
\caption{Schematic of the JENSA system as installed at NSCL. The major gas flow during operation is indicated with green arrows. The gas flow through the nozzle and hence the jet density is adjustable by needle valve \textcolor{c12}{V$_\text{LO}$} regulating the flow through the loop (violet arrows). Less than 10\% of the gas is caught by the outer receiver (orange arrows). Regions of similar pressure are grouped by light gray shadows. The different kinds of pumps are described in detail in the text.}
\caption{Spectrum from an activity standard of $^{228}$Th measured by front strip \#48 of the BB15 detector. Gaussian functions (black curves) are fitted to the detected$\alpha$ counts per channel (blue histogram), including peaks at (from left to right) 5340.36, 5423.15 (both from the decay of $^{228}$Th), 5685.37 ($^{224}$Ra), 6050.78 ($^{212}$Bi), 6288.33 ($^{220}$Rn), 6778.3 ($^{216}$Po) and 8784.86\,keV ($^{212}$Po). Empty channels between 7.2 and 8.3\,MeV are skipped. With similar spectra for each of the 64~front strips, the region of the energy loss measured with an activity standard of$^{241}$Am (\autoref{fig:americium}) has been calibrated by interpolating the known peaks.}
\caption{Spectrum from an activity standard of $^{241}$Am measured by front strip \#40 in coincidence with back strip\#3. The$\alpha$ particles were measured without (black histogram) and with a jet at an input pressure of 2.514\,MPa (purple histogram). Thick curves are double Gaussian functions fitted to the histograms using the nonlinear least-squares Marquant-Levenberg algorithm~\cite{Levenberg1944QAM,Marquardt1963JSIAM}. The high-energy peaks correspond to $\alpha$ particles leaving the collimator in front of the source with the full energy of 5485.56-keV. The energy difference of the fitted maximums (dashed vertical lines) correspond to an energy loss of $57.5\pm0.8$\,keV in the jet. The low-energy shoulders of both peaks result from energy losses at the edge of the collimator bore-hole and from another$^{241}$Am decay emitting $\alpha$ particles with 5442.8 keV. Note that similar spectra have been analyzed for a total of 256 pixels $\times$ 6~nozzles $\times$ 3 input pressures. The determined energy losses are shown in \autoref{fig:eloss}.}
\caption{Energy loss of 5486-keV $\alpha$ particles crossing through the upper, middle and lower (horizontal strip number 3, 2 and 1) part of the jet at compressor discharge pressures of 1.136\,MPa (orange), 1.825\,MPa (green) and 2.514\,MPa (purple) of six different nozzles (\autoref{tab:nozzles}). The upper part indicates the first 4\,mm of the jet, the middle part 4 to 8\,mm and the lower part 8 to 12\,mm. The energy loss for all jet segments at a single pressure is shown in each panel of the left three columns. The right three columns show the energy loss for a single jet segment at three different pressures. Horizontal strip number 4 is partly shadowed by the nozzle (empty squares at vertical strip numbers 37 to 44 in the left panels) and hence not shown in the right panels. Solid black lines lines are Gaussian functions fitted to the measured energy losses per vertical strip number using the nonlinear least-squares Marquardt-Levenberg algorithm~\cite{Levenberg1944QAM,Marquardt1963JSIAM}. Note that the first three rows of the right three columns have been measured with Nozzle A, rows 4 to 6 with nozzle~B and so on. The ambient pressure in the surrounding vacuum chamber is given in \autoref{tab:discharge_flow}. The negative energy-loss artifacts for the measurements of nozzle~E and F are caused by the noisy front strip \#53, that is not influencing the fit region of the peak.}
\caption{Averaged SSIM and PSNR value on synthesized images with their parameter number. {\color{red}Red} indicates the best and {\color{blue}blue} indicates the second best performance. }
\caption{ \footnotesize Timeplot delay measurements for 10s.Terms: C=Link Capacity,. Terms: \textcolor[rgb]{0,0,1}{A1=Actuator1 (Blue)}, \textcolor[rgb]{1,0,0}{A2=Actuator 2 (Red)}, BE=Best-effort queue, ST=Scheduled Traffic queue. a) Same priority blocking for C=1 Gbps b) Lower priority blocking for C=1 Gbps c) Lower priority blocking for C=100 Mbps d) Using a TAS for C=1 Gbps}
\caption{\footnotesize Timeplot delay measurements for 10s. Link Capacity 100 Mbps. Terms: \textcolor[rgb]{0,0,1}{A1=Actuator1 (Blue)}, \textcolor[rgb]{1,0,0}{A2=Actuator 2 (Red)}, BE=Best-effort queue, ST=Scheduled Traffic queue. a) Same priority blocking b) Same priority blocking with network load c) Lower priority blocking d) Lower priority blocking with network load e) Using a TAS f) Using a TAS with network load.}
\caption{\footnotesize Timeplot delay measurements for 10s. Link Capacity 1 Gbps. Terms: \textcolor[rgb]{0,0,1}{A1=Actuator1 (Blue)}, \textcolor[rgb]{1,0,0}{A2=Actuator 2 (Red)}, BE=Best-effort queue, ST=Scheduled Traffic queue. a) Same priority blocking b) Same priority blocking with network load c) Lower priority blocking d) Lower priority blocking with network load e) Using a TAS f) Using a TAS with network load.}
\caption{Quantitative results for our proposed method (ConnNet) compared to other recent approaches. To illustrate the effect of including global relations, we use $CONN$ and $CONN+$ to denote ConnNet without and with global relations, respectively. We exclude results for the test results on the MSRA-B dataset for the RFCN and the DHSNet, as they were included in the respective training datasets following~\cite{li2017instance}. We report maximum F-measure (larger is better) and highlight the best three results for each dataset in the colors {\bf\textcolor[rgb]{0.8,0.4,0.4}{orange}}, {\bf\textcolor{blue}{blue}} and {\bf\textcolor{darkgreen}{green}}, respectively. {DSS$^\dagger$} represents a recently improved version of DSS that utilizes a pretrained ResNet-101.}
\caption{Example QA pairs obtained from the original movie plot and the paraphrased plot. The relevant spans needed for answering the corresponding question are highlighted in {\color{blue}blue} and {\color{red}red} with the respective question numbers. Note that the span highlighting shown here is for illustrative purposes only and is not available in the dataset.}
\caption{A hybrid word-character models for number generation, in which numbers are highlighted in {\color{blue} blue} color. For simplicity, we skip the memory component. }
\caption{An example to illustrate non-syntactic tokens: \textcolor{red}{sentiment emoticon}, \textcolor{orange}{retweet marker and its following at-mention}, \textcolor{gray}{topical hashtag}, \textcolor{blue}{referential URL}, and \textcolor{cyan}{truncated word}. This is a concatenation of three real tweets. }
\caption{({\color{blue}a}) Schematic representation of the Deutsch's circuit. ({\color{blue}b}) Circuit and schematic representation of two-level system associated to alternative approach presented in this paper.}
\caption{Vectorization examples from (1) 444.namd benchmark and (2) BT benchmark in C like pseudocode (a) scalar code (1)(b) and (1)(c) show LLVM vectorized version and \projectName{} vectorized version for 444.namd respectively (2)(b) shows \projectName{} vectorized version for BT; Vectorized code is shown in \textcolor{blue}{blue}, non-vectorizable code that is packed into vectors is shown in \textcolor{maroon}{maroon} and any unpackings of vectors are shown in \textcolor{darkgreen}{dark green}. Unpackings are shown as indexing into the proper lane of the relevant vector value (e.g., \texttt{V1[0]} denotes extracting the $0^{\textrm{th}}$ lane from vector \texttt{V1}).}
\caption{{\it Left panel}: The differential detection rate of gaseous halos as a function of galaxy stellar mass assuming the maximum radius is the virial radius. The adopted SMF is only for star-forming galaxies \citep{Tomczak:2014aa}. \red{{\it Right panel}: The cumulative detection rate of gaseous halos.}}
\caption{Inter-beat (RR) intervals were computed as distance between successive peaks (marked in \textcolor{red}{red}) in the ECG wave to calculate HRV.}
\caption{Face localized points marked in \textcolor{green}{green} and a subset of features used to calculate face AUs marked as \textcolor{cyan}{cyan} lines. Height (H) and width (W) of the face were used to normalize the features. Subject's consent to use her face for publication is marked in the dataset.}
\caption{\label{fig:coef} Coefficient values for coarse features in the paper acceptance classification, for {\color{blue}{ICLR}} and {\color{red}{arXiv}}.}
\caption{SegFinNet with visual attention mechanism for two different input latents, one per row: (a) Focused region from Visual attention module (section \ref{sec:wheretolook}); (b) Original latents overlaid with a heat map showing the probability of occurrence of friction ridges (from \textcolor{red}{high} to \textcolor{blue}{low}); (c) Binary mask (boundary marked in red) used in subsequent modules: enhancement, feature extraction, and matching.}
\caption{Different groundtruths for two latents in NIST SD27. The groundtruth croppings shown in \textcolor{red}{red}, \textcolor{OliveGreen}{green}, and \textcolor{blue}{blue} are used in Cao \etal~\cite{cao2014segmentation}, Ruangsakul \etal~\cite{ruangsakul2015latent}, and Zhu \etal~\cite{zhu2017latent}, respectively.}
\caption{Left: Overview of the reasoning in a standard line of argument \cite{Fermi56, Zemansky68, Pauli73, Adkins83}. Eventually, Carnot's theorem establishes the relation between the previously introduced empirical and absolute temperature. The color of the boxes indicates whether they are \textcolor{red}{definitions}, \textcolor{blue}{assumptions and postulates}, or \textcolor{green}{derived implications} in each setting. Right: Overview of the reasoning presented in this paper.}
\caption{The main line of argument as presented in \cite{LY99} (left) and in the framework used in this paper (right). The color code is the same as in Figure \ref{fig:overview}: \textcolor{red}{definitions}, \textcolor{blue}{assumptions and postulates}, and \textcolor{green}{derived implications}. Left: Defining an order relation $\succ$ and postulating several properties of it, Lieb and Yngvason are able to prove a theorem they call the ``entropy principle''. From there, the definition of entropy is obtained and it is possible to show that the second law holds as well as Carnot's theorem. Other approaches based on a second law similar to Carath\'eodoy's\cite{Caratheodory09} proceed in a similar way. Right: Overview of the reasoning presented in this paper and a follow-up paper \cite{Kammerlander18}. After introducing absolute temperature as presented before we will be able to define the absolute temperature of heat flows between arbitrary systems. Based on this, the usual definition of thermodynamic entropy can be used and theorems such as Clausius' theorem or the entropy principle can be proved. }
\caption{(Color online) \textbf{LGR driving of an NV ensemble}. \textbf{(a)} NV electron spin resonance spectrum (\textcolor{blue}{\textbf{---}}) under application of bias field $B_0$. The bias field allows individual addressing of all eight NV resonances, arising from the combination of the two allowed magnetic dipole transitions with the four possible NV orientations. The NV hyperfine structure is obscured by MW power broadening and the contrast variation between the NV resonances is attributed primarily to the $S_{11}$ line-shape. % and polarization-induced variation in the dipole coupling between each NV class and the optical excitation The $S_{11}$ parameter is shown before (\textcolor{red}{\textbf{-\,-\,-}}) and after (\textcolor{dolla-bill}{\textbf{-\,-\,-}}) shifting the LGR resonant frequency $f_0$ to the target NV resonance. Arrows indicate corresponding y axes. \textbf{(b)} Typical data depicting Rabi oscillations under MW excitation at the target NV resonance frequency indicated in (a). Data (\textcolor{navyblue}{$\mathbf{\circ}$}) is fit (\textcolor{blue}{\textbf{---}}) to an exponentially decaying sinusoid. %(See appendix \ref{app}). }
\caption{(Color online) \textbf{$\boldsymbol{B_1}$ field uniformity of LGR composite device.} \textbf{(a)} An NV-containing 4.5 mm $\times$ 4.5 mm diamond plate is placed in the LGR central loop, and the Rabi frequency is measured where indicated (\textcolor{deepmagenta}{\textbullet},\textcolor{black}{\textbullet},\textcolor{red}{\textbullet},\textcolor{blue}{\textbullet}) to characterize $B_1$. \textbf{(b)} Simulations suggest the $B_1$ field distribution should be approximately radially symmetric, with the leading order deviation resulting from the exciter antenna. Dashed lines indicate the 32 mm$^2$ and 11 mm$^2$ areas within which the $B_1$ field uniformity is evaluated. \textbf{(c)} $B_1$ field measurements (\textcolor{deepmagenta}{$\circ$},\textcolor{black}{$\circ$},\textcolor{red}{$\circ$},\textcolor{blue}{$\circ$}) at the points depicted in (a) and simulations (\textcolor{deepmagenta}{\textbf{--}},\textcolor{black}{\textbf{--}},\textcolor{red}{\textbf{--}},\textcolor{blue}{\textbf{--}}) along each locus of points are in good agreement. Error bars indicate 1-sigma uncertainty for the $B_1$ measurement. Dashed lines indicate the radial boundaries of the 32 mm$^2$ and 11 mm$^2$ areas over which $B_1$ field uniformity is evaluated. The measured $B_1$ uniformity is given for each area.}
\caption{Optic Nerve Head (ONH) region imaged with commercialized instruments and Laser Doppler holography. (a) Scanning laser ophthalmoscope (Spectralis, Heidelberg). (b) Adaptive optics flood illumination (rtx1, Imagine Eyes). (c) Power Doppler images $M_{0^{+}}$ calculated from holograms recorded at $f_{\rm S}=39 \, \rm kHz$; $S(\omega)$ is integrated over $[f_{\rm 1}, f_{\rm 2}]$ = 4-19.5 kHz. Multiple power Doppler images $M_{0^{+}}$ are averaged over a total time of $80 \, \rm ms$ (see \textcolor{blue}{Visualization 1} for blood flow movie). The peri-papillary crescent is visible on the edge of the ONH. (d) Asymmetry of the DPSD $M_{0^{-}}$ (averaged over the same period of time) illustrating the resultant flow direction with respect to the optical axis.}
\caption{An artery intertwined with a vein is imaged with commercialized instruments and Laser Doppler holography.(a) Scanning laser ophthalmoscope (Spectralis, Heidelberg). (b) Adaptive optics flood illumination (rtx1, Imagine Eyes). (c) Power Doppler images $M_{0^{+}}$ calculated from holograms recorded at $f_{\rm S}=39 \, \rm kHz$; $S(\omega)$ is integrated over $[f_{\rm 1}, f_{\rm 2}]$ = 7-19.5 kHz. Multiple power Doppler images $M_{0^{+}}$ are averaged over a total time of $0.66 \, \rm s$ (see \textcolor{blue}{Visualization 2} for blood flow movie). (d) Asymmetry of the DPSD $M_{0^{-}}$ (averaged over the same period of time) illustrating the resultant flow direction with respect to the optical axis. }
\caption{Requirements in terms of sampling frequency for laser Doppler holography in the central retinal artery. For each sampling frequency, power Doppler images are spatially averaged over the depicted ROIs on the left hand side and the result is displayed in the associated plot. The STFT parameters used for each acquisition are displayed on the left and have been chosen to have $\sigma_{\rm win} \approx 13 \, \rm ms$ for all acquisitions. (a) $f_{\rm S}=10 \, \rm kHz$, (b) $f_{\rm S}=20 \, \rm kHz$, (c) $f_{\rm S}=39 \, \rm kHz$, (d) $f_{\rm S}=75 \, \rm kHz$. \textcolor{blue}{Visualization 3} shows PDIs as a function of the frequency for the acquisition with $f_{\rm S}=75 \, \rm kHz$. }
\caption{Additional \aastex\symbols}
\caption{Building a transplant net. We propose a theoretical solution to incrementally merging category modules from teacher nets into a transplant (student) net with a few or without sample annotations. The transplant net has an interpretable, modular structure. A category module, \emph{e.g.} a cat module, provides cat features to different task modules. A task module, \emph{e.g.} a segmentation module, serves for various categories. We show two typical operations to learn transplant nets\textcolor{red}{{\protect\footnotemark[1]}}. Blue ellipses show modules in teacher nets used for transplant. Red ellipses indicate new modules added to the transplant net. Unrelated adapters in each step are omitted for clarity.}
\caption{\textbf{Model Overview:} The \textcolor{red}{red} dotted lines represent the loss functions, the downward diagonal arrows ($\protect\shortarrow{7}$) represent the decoding operation and the upwards diagonal arrows ($\protect\shortarrow{1}$) represent the encoding operation. The proposed model first encodes the input image ($\mathbf{x}$) to a parameterized latent representation ($\mathbf{z}_v$) and reconstructs it back using the generator ($\mathbf{x}_g$ and $\mathbf{x}_v$ corresponding to latent representations $\mathbf{z}_g$ and $\mathbf{z}_v$ respectively). The discriminator also consists of an auto-encoder which first encodes the inputs to a latent representation ($\mathbf{z}_d$, $\mathbf{z}'_g$ and $\mathbf{z}'_v$ corresponding to inputs $\mathbf{x}$, $\mathbf{x}_g$ and $\mathbf{x}_v$ respectively), and then reconstructs them back ($\mathbf{x}'_d$, $\mathbf{x}'_g$ and $\mathbf{x}'_v$). The generator and discriminator are trained using the back-propagated error signal from the losses computed to match the actual data, the latent representation and the loss distributions for the real and fake samples.}
\caption{Quantitative comparison on the CIFAR-10 dataset in terms of Inception score. The best and the second best performances are shown in \textcolor{red}{red} and \textcolor{blue}{blue} respectively. BEGAN* denotes \cite{berthelot2017began} with the modified equilibrium approach and LSM stands for the Learned Similarity Metric. All the reported performances are for unsupervised cases except the bottom two, which use label information or real images, respectively. }
\caption{Additional \aastex\symbols}
\caption{For the simulations with adiabatic electrons (left column) and kinetic electrons (right column), from top to bottom: calculated potential in W7-X at the toroidal planes $\zeta=0, 15, 36$ and $54^\circ$. \blue{Note the different color scales on the left and right plots, employed to appreciate the changes in the shape of $\Phi_1$ between considering adiabatic or kinetic electrons.}}
\caption{Maximum difference of the potential $\Delta\Phi_1=\Phi_{1}^{\text{max}}-\Phi_{1}^{\text{min}}$ on each of the simulated flux surfaces normalized to the ion temperature $T_{i}$ as a function of the normalized effective radius $r/a$ for the calculations with adiabatic electrons (circles connected with blue segments) and kinetic electrons (squares connected with red segments). \blue{The dashed vertical line indicates the radial position where $E_{r}=0$. On the left and right of this line the input radial electric field is positive and negative, respectively.}}
\caption{\label{Fig:1}Storage of a single photon in the electronic state of a single atom confined inside an optical resonator. (a) The photon wave packet propagates along a transmission line and impinges onto a cavity mirror. (b) The single photon is absorbed by the cavity, which drives the atomic transition $|g\rangle \to |e\rangle$. An additional laser couples to the atomic transition $|r\rangle \to |e\rangle$. The dynamics of storage is tailored by optimizing the functional dependence of the laser amplitude on time, $\Omega(t)$: Ideally, the atom undergoes a Raman transition to the final state $|r\rangle$ and the photon is stored. We analyse the storage \red{efficiency} including the spontaneous decay with rate $\gamma$ of the excited state and photon absorption or scattering at the cavity mirrors via an incoherent process at rate $\kbad$. Further parameters are defined in the text.}
\caption{\label{fig:4}Storage \red{efficiency} $\eta$ at $t=t_2$ as a function of the coherence time of the single-photon pulse $\Tc$ (in units of $(\gamma C)^{-1}$). The legenda indicates the pulses used in the numerical integration of Eq. \eqref{eq:mastereq}. The parameters are $(g,\kappa)=(4.9,2.42)\MHz$, the lines labeled ``with losses'' refer to the \red{efficiency} of the process when $(\gamma,\kbad)=(3.03,0.33)\MHz$, otherwise $\gamma=\kbad=0$; $t_2=-t_1=6\Tc$. The other parameters are given in Fig. \ref{fig:2}.}
\caption{\label{fig:6}Minimum photon coherence time $\Tc^\mathrm{min}$ as a function of $g$ (in units of $\kappa$). The coherence time $\Tc^\mathrm{min}$ is the lower bound to the coherence time of photons which can be stored with \red{efficiency} (a) $\eta_\mathrm{tr}=0.99$ and (b) $\eta_\mathrm{tr}=2/3$ for $\gamma=\kbad=\Delta=0$. The vertical dotted line shows the value $g=\kappa=2.42\MHz$. The data in the region $g\ll\kappa$ and $g\gg\kappa$ have been fitted with the functions $f_1(g) = a \kappa/g^2$ and $f_2(g) = a'/ \kappa$, respectively.}
\caption{\label{fig:comparisonDelta}(a) Storage \red{efficiency}, (b) probability of photon reflection, Eq.~\eqref{eq:pr}, and (c) probability of spontaneous decay, Eq.~\eqref{eq:ps}, as a function of the single photon detuning $\Delta$ and at time $t_2$. The parameters are $(g,\kappa,\gamma)=(4.9,2.42,3.03)\MHz$, $\Tc=0.5\unit{\mu s}$. The input photon $\Ein(t)$ is defined as in Eq.~\eqref{eq:Einhypsec}. See Fig. \ref{fig:2} for further details.}
\caption{\textit{Ab initio} equilibrium molecular structures for \ce{SiH3CN} (top) and \ce{SiH3NC} (bottom). Values in {\color{red}red} were obtained with a cc-pVQZ basis, while the calculated values in black include core-valence basis functions (cc-pwCVQZ). Bond lengths are given in \AA, and angles in degrees. See SI for analogous calculations of \ce{SiH3CN}.}
\caption{Complexity-accuracy tradeoff of various DNN architectures quantized using different techniques. Complexity is reported in number of bit operations as explained in the text. Number of bits is reported as (weights,activations) and model size is calculated as the sum of parameters sizes in bits. Accuracy is top-1 accuracy on ImageNet. For each DNN architecture, rows are sorted in increasing order complexity. Note that for the same architecture with same bitwidths UNIQ may have a different complexity and model size due to fact that, unlike the others, we also quantize the first and the last layers. Compared methods: \textcolor{color_xnor}{XNOR} \cite{rastegari2016xnor}, \textcolor{color_qnn}{QNN} \cite{hubara2016quantized}, \textcolor{color_iqn}{IQN} \cite{zhou2017incremental}, \textcolor{color_appr}{Apprentice} \cite{mishra2017apprentice}, \textcolor{color_dist}{Distillation} \cite{polino2018model}, \textcolor{color_qsm}{QSM} \cite{DBLP:journals/corr/abs-1803-08607}, \textcolor{color_mlq}{MLQ} \cite{AAAI1816479}. }
\caption{Spectrum of WASP-121b at different orbital phases, $\alpha=0^{\circ}$ being the secondary eclipse. Emission features are expected for $-120^{\circ}<\alpha<60^{\circ}$ whereas absorption features are seen for $\alpha<-120^{\circ}$ and $\alpha>60^{\circ}$. The spectra at phases $-120^{\circ}$ and $-60^{\circ}$ \textcolor[rgb]{0.984314,0.00784314,0.027451}{are} well represented by an area weighted sum of the dayside and the nightside spectrum (dotted line) but the $120^{\circ}$ and $60^{\circ}$ spectra are not.}
\caption{Examples of human provided (\textcolor{Green}{green}) and captioner generated relative descriptions (\textcolor{NavyBlue}{blue}). While generated relative captions don't resemble human annotations in most cases, they can nonetheless capture the main visual differences between the target image and the reference image.}
\caption{Example where encoding the distance of the object furthest away (solid \textcolor{darkgreen}{green}) is better than that of the one closest to the image centre (dashed \textcolor{darkred}{red}). The IC model assumes that only one person is in the middle in the former case, and infers that many people may be gathered around a table in the latter.}
\caption{Reynolds stresses for impulses at $z^+=30$ (top row), $z^+=500$ (middle row), and $z=-0.5$ (bottom row) with forcing along $x$ (left columns), $y$ (middle columns), and $z$ (right columns) at times $tU_{CL}/h$ from $5.7$ to $42.7$ in steps of $2.84$; $I_{11}$ with scale on the bottom $x$-axis: blue to red for increasing time, {\color{redBlue}\full}; $I_{13}$ with scale on the top $x$-axis: black to green for increasing time, {\color{midGreen}\dashed}~ for structures with $\delta_h \leq 0.5$ and {\color{midGreen}\dotted}~ otherwise. The average locations of the peaks of $I_{11}$ ({\color{cyan}\dashdot}) and $I_{13}$ ({\color{yellow}\dashdot})are also shown. All curves are truncated to ignore the top half of the channel ($z>1$). Insets show near-wall streamwise energy: $I_{11}$/max($I_{11}$) on the $x$-axis against $z^+$ on the $y$-axis in log scale; note that $z^+$ here is not scaled by $\delta_h$.\label{fig:stress-instant-multiple}}
\caption{Evolution of total ({\color{red} \dashed}) and streamwise ({\color{blue} \dotted}) kinetic energy density for impulses introduced at different locations: $z^+ \approx 30$ (left), $z^+ \approx 500$ (middle), and $z = -0.5$ (right). For each location, the plots show energy due to impulses along streamwise ($+$), spanwise ($\times$), and wall-normal ($\triangledown$) directions. Markers correspond to times where coherent structures are plotted in \cref{fig:swirl-lapse-full}.\label{fig:energy-multiple}}
\caption{\cred{Top: kinetics of voltage, gating variable (left axis) and cell shortening with active strain function (right axis) plotted against time. Bottom: Variations of the dynamics of the coupled thermoelectric model according to temperature. The variations of the active strain and myocyte shortening coincide and therefore the latter are not shown.}}
\caption{\cred{Example of approximate displacement field on the deformed domain and a coarse hexahedral mesh in the undeformed configuration (left); and error history for an accuracy test of the mixed formulation for hyperelasticity (right) generated using a lowest-order discretisation.}}
\caption{Samples of the approximate solutions (potential, activation, displacement magnitude, and solid pressure) shown on the deformed domain at \cred{$t=50,150,450$\,ms (left, middle, and right panels)}. For this test we have used $T=39^\circ$C. }
\caption{\cred{Evolution of main variables measured on the point (5,2.5,2.5) and up to $t=720$\,ms, computed at temperature$T=39^\circ$C.} }
\caption{\cred{Potential difference between case I (uniform temperature at $T=37^\circ$C) and two different gradient distributions in the sheetlet (case II - left) and radial directions (case III - right), plotted on the reference domain at times $t=350$\,ms (top panels),$t=450$\,ms (middle row), and$t = 600$\,ms (bottom panels)}.}
\caption{\cred{Ellipsoidal fibre distribution, collagen normal-sheetlet, and cross-fibre directions generated with a rule-based algorithm and setting} $\theta_{\text{epi}}=-50^\circ$, $\theta_{\text{endo}}=60^\circ$.}
\caption{Propagation of the transmembrane potential plotted on the deformed domain, using a constant temperature (top panels) and a cold spot (bottom). \cred{Snapshots shown at 200,300,400,500\,ms after the S2 stimulus.}}
\caption{\cred{Sample approximate Kirchhoff stress, displacement, pressure, and myocyte shortening at end diastole, $t = $470\,ms (top) and$t = $610\,ms (bottom).}}
\caption{\cred{Propagation of the transmembrane potential plotted on the deformed domain, using a higher temperature throughout the domain and a thiner ventricular geometry. Snapshots shown at 100,200,\ldots,800\,ms after the S2 stimulus.}}
\caption{(Color online) $\Delta n_z$ (a,b) as a function of layer index $z$ for different values of the Hubbard interaction $U$. (c,d) $\Delta n_z$ for the LIL (blue curve) and for the CIL (red curve) as a function of Hubbard interaction $U$. Dashed green lines in (e) show results of the fit for the CIL $\Delta n_{\rm CIL}=A_1 \exp(-A_0 U)$ and for the LIL $\Delta n_{\rm LIL}=B_2+B_1 \exp(-B_0 U)$, with fit parameters $A_0\sim B_0=0.301$, $A_1=-0.139$, $B_1=0.027$, and $B_2=0.090$. In (d) we additionally plot the curves %$0.04 V(U)$ {\color{green2}$0.04*V$ versus $U$} (green). In the inset of (d) we plot the bias voltage as a function of interaction $U$ for fixed value of $\Delta \mu=0.5$ (see details in text). (e,f) double occupancy $d_z= \langle n_{z,{\bf r},\uparrow} n_{z, {\bf r} \downarrow}\rangle$ for the CIL (red curve) and for the CML(indigo curve) as a function of Hubbard interaction $U$. The results in (a,c,e) are obtained with $\Delta \mu =2$ and $t_{lc}=t_{rc}=0.2$, while the ones in (b,d,f) with $\Delta \mu =0.5$ and $t_{lc}=t_{rc}=1$. Other parameters are the same as in Fig.~\ref{Fig:N_U4U8tlc0p2tlc1}. }
\caption{Power spectral analysis of pressure $p/(\frac{1}{2}\rho_\infty u_\infty^2)$ on the aft wall for the baseline spanwise-periodic and finite-span cavity flows at $M_\infty=0.6$ and 1.4. The probe location is $[x,y,z]/D=[6,-0.5,0]$. Spanwise-periodic case: {\color{blue}{\bf --} (blue)}, and finite-span case: {\color{red}{\bf --} (red)} with shaded uncertainty bounds representing 95\% confidence. The predicted Rossiter mode frequencies using Eq.~(\ref{RossiterFormula}) are denoted by dashed lines.}
\caption{Contours of time-averaged streamwise velocity $\bar u/u_\infty$ (dashed lines for negative values) with an increment of $\Delta \bar u/u_\infty=0.1$, and quiver of time-averaged velocities [$\bar v$, $\bar w$]/$u_\infty$ colored by $\bar v/u_\infty$ on $y$-$z$ plane at $x/D=5.5$ for the baseline flows. The dot-dashed black lines indicate the cavity midspan, while the dashed blue lines indicate the periodic boundaries. The red dot {\color{red}$\bullet$} highlights a contour line of $\bar u=0$. }
\caption{The power spectral analysis of pressure $p/(\frac{1}{2}\rho_\infty u_\infty^2)$ on the aft wall for the controlled spanwise-periodic and finite-span cavity flows at $M_\infty=0.6$ and 1.4. The probe location is at $[x,y,z]/D=[6,-0.5,0]$. Spanwise-periodic case: {\color{blue}{\bf --} (blue)}, and finite-span case: {\color{red}{\bf --} (red)} with shaded uncertainty bounds representing 95\% confidence. The predicted Rossiter mode frequencies using Eq.~(\ref{RossiterFormula}) are denoted by dashed lines. The power spectra for baseline cases are shown in figure \ref{fig:p_base_rm}.}
\caption{Contours of time-averaged streamwise velocity $\bar u/u_\infty$ (dashed lines for negative values) with an increment of $\Delta \bar u/u_\infty=0.1$, and quiver of time-averaged velocities [$\bar v$, $\bar w$]/$u_\infty$ colored by $\bar v/u_\infty$ on $y$-$z$ plane at $x/D=5.5$ for the controlled flows. The dot-dashed lines indicate the cavity midspan, while the dashed blue line indicates the periodic boundaries. The red dot {\color{red}$\bullet$} highlights a contour line of $\bar u=0$. The same plots for baseline flows are shown in figure \ref{fig:slice55_base}. }
\caption{Predictions from a vanilla \seqseq{} model, illustrating {\color{red}unsupported facts}, {\color{blue}missing facts} and {\color{darkgreen}repeated facts}. The last two rows show inputs we composed to demonstrate that the models memorize entity-fact pairs.}
\caption{The OC-SVM pipeline. The image is first cropped (a) and then superpixeled (b). The OC-SVM is trained ($\gamma = 2^{-20}$, $\nu = 0.3$) using the superpixels considered \textcolor{red}{background} in (c), and those with an unknown label (uncoloured) have their class predicted, giving a final predicted object mask (d).}
\caption{The SBBM pipeline. The image is cropped (a) and superpixeled (b). A model ($\delta = 0.5$) is created for each \textcolor{red}{superpixel} in (c) outside the BB, and the other superpixels (uncoloured) are compared to them ($\eta = 0.8$) and classified (d).}
\caption{The alpha matting pipeline. (a) The \textcolor{cyan}{original}, \textcolor{red}{expanded} ($\rho^+ = 1.7$), and \textcolor{ForestGreen}{contracted} ($\rho^- = 0.9$) BBs. The calculated alpha matte (b) and the mask (c) corresponding to a threshold of $\tau = 0.80$.}
\caption{Alpha matting: (a) Original image. (b) Cross validated ($\rho^- = 0.8, \rho^+ = 1.2, \tau = 0.85, \lambda = 10^{-2}$) and (c) best possible segmentations ($\rho^- = 0.6, \rho^+ = 1.1, \tau = 0.85, \lambda = 10^{-2}$), with labelled \textcolor{cyan}{original}, \textcolor{red}{expanded}, and \textcolor{ForestGreen}{contracted} BBs.}
\caption{Distribution of the \textcolor{blue}{optimal} and \textcolor{ForestGreen}{cross-validated} alpha matting parameters for the $\phi_{all}$ measure. (a) $\rho^-$, (b) $\rho^+$, (c) $\tau$, and (d) $\log_{10}(\lambda)$.}
\caption{Printed sample with 9 defects 4~mm by 4~mm in size, each 0.5~mm thick, spaced 10~mm apart in both directions. Sample was 70~mm by 56~mm and 8~mm thick. \textcolor{blue}{Blue} numbers above the defects indicate depth below the inspected surface and \textcolor{red}{red} horizontal lines with numbers indicate the numbered profiles where data was extracted from measurements. \label{fig:printedsample1}}
\caption{Printed sample with square defects from 2~mm to 12~mm in width, each either 0.5~mm or 1~mm thick. Sample was 70~mm by 56~mm and 6~mm thick. \textcolor{blue}{Blue} numbers above the defects indicate depth below the inspected surface and \textcolor{red}{red} horizontal lines with numbers indicate the numbered profiles where data was extracted from measurements. \label{fig:printedsample2}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, at 1~frame per second, for 200~seconds, for PLA samples shown in Figure~\ref{fig:printedsample2}. \textcolor{black}{Black} lines represent data extracted from a sample with defects that were 0.5~mm thick. \textcolor{red}{Red} and \textcolor{blue}{blue} lines represent data extracted from vertically and horizontally printed samples with defects 1.0~mm thick. Data is presented for profiles 1, 4, 6, and 8. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phaseprintdir}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, at 1~frame per second, for 200~seconds, for samples shown in Figure~\ref{fig:printedsample2}. \textcolor{black}{Black} and \textcolor{red}{red} lines represent data extracted from PLA and ABS samples with defects that were 0.5~mm thick. Data is presented for profiles 1, 4, 6, and 8. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsqdiff}}
\caption{Lock-in phase signals from measurements conducted at different lock-in frequencies, at 1~frame per second, for 200~seconds, for samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black} and \textcolor{red}{red} lines represent data extracted from PLA and ABS samples with defects that were 0.5~mm thick. Data is presented for profile 1. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematfreqprof1}}
\caption{Lock-in phase signals from measurements conducted at different lock-in frequencies, at 1~frame per second, for 200~seconds, for samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black} and \textcolor{red}{red} lines represent data extracted from PLA and ABS samples with defects that were 0.5~mm thick. Data is presented for profile 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematfreqprof2}}
\caption{Lock-in phase signals from measurements conducted at different lock-in frequencies, at 1~frame per second, for 200~seconds, for samples shown in Figure~\ref{fig:printedsample1}, for PLA~(top) and ABS~(bottom) samples with defects that were 0.5~mm thick. Signals from centroid of the 9 defects were used to obtain phase signals with respect to depth, at lock-in frequencies of 0.01~Hz (\textcolor{black}{$\times$, solid}), 0.025~Hz (\textcolor{red}{$\times$, dashed}), and 0.05~Hz (\textcolor{blue}{$\times$, dotted}). \label{fig:depthresolution}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for PLA samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black}, \textcolor{red}{red}, \textcolor{blue}{blue}, \textcolor{OliveGreen}{green}, and \textcolor{Dandelion}{orange} lines represent data extracted from using SC7500 at 1 frame per second, A310 at 2 frames per second, Ti55FT at 0.5 frames per second, Seek Thermal CompactXR at 1 frame per second and Gobi 640 at 1 frame per second respectively. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematcam}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data for low, medium and high power. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpow}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data with baseline correction for low, medium and high power. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpowoffset}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}, using Ti55FT. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data for low, medium and high power. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpowTi55}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}, using Ti55FT. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data with baseline correction for low, medium and high power. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpowoffsetTi55}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data for low, medium and high power. Corresponding dashed lines were obtained from constant heating (not modulated) such that the mean heating would be the same, and processed using the same lock-in frequency. These lines were shifted upwards by 0.3 (\textcolor{black}{black}), 0.48 (\textcolor{red}{red}), and 0.50 (\textcolor{blue}{blue}) respectively to compare with results from modulated heating. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpowjustify}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}, using Ti55FT. \textcolor{black}{Black}, \textcolor{red}{red} and \textcolor{blue}{blue} lines represent data for low, medium and high power. Corresponding dashed lines were obtained from constant heating (not modulated) such that the mean heating would be the same, and processed using the same lock-in frequency. These lines were shifted upwards by 0.24 (\textcolor{black}{black}), 0.41 (\textcolor{red}{red}), and 0.42 (\textcolor{blue}{blue}) respectively to compare with results from modulated heating. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematpowTi55justify}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black} and \textcolor{red}{red} lines represent data for synchronous (interpolated images) and asynchronous Lock-In Thermography. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsync}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. \textcolor{black}{Black} and \textcolor{red}{red} lines represent data with baseline correction for synchronous (interpolated images) and asynchronous Lock-In Thermography. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsyncoffset}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. Full sampling (\textcolor{black}{Black}) was compared with subsampling of every 2 (\textcolor{red}{red}), 4 (\textcolor{blue}{blue}), 8 (\textcolor{black}{black dashed}), 10 (\textcolor{red}{red dashed}) and 20 (\textcolor{blue}{blue dashed}) frames from the full data. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsubsamp}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. Full sampling (\textcolor{black}{Black}) was compared with subsampling of every 2 (\textcolor{red}{red}), 4 (\textcolor{blue}{blue}), 8 (\textcolor{black}{black dashed}), 10 (\textcolor{red}{red dashed}) and 20 (\textcolor{blue}{blue dashed}) frames from the full data, with baseline correction of all data. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsubsampoffset}}
\caption{Lock-in phase signals from measurements conducted at 0.01~Hz lock-in frequency, for 200~seconds, for ABS samples shown in Figure~\ref{fig:printedsample1}. Full sampling (\textcolor{black}{Black}) was compared with subsampling of 100 images with first frame equal to 0~(\textcolor{red}{red}), 10~(\textcolor{blue}{blue}), 20~(\textcolor{black}{black dashed}), 30~(\textcolor{red}{red dashed}), 40~(\textcolor{blue}{blue dashed}), 50~(\textcolor{black}{black dashed dotted}), 60~(\textcolor{red}{red dashed dotted}), 70~(\textcolor{blue}{blue dashed dotted}), 80~(\textcolor{black}{black dotted}), 90~(\textcolor{red}{red dotted}), 100~(\textcolor{blue}{blue dotted}) frames from the first frame, with baseline correction of all data. Data is presented for profiles 1 and 2. \textcolor{magenta}{Magenta} boxes represent the locations of known defects, with numbers indicating the depth of defects. \label{fig:phasematsubsampstarttime}}
\caption{The contact angle is measured from photographic images of droplets formed on the substrate using MATLAB\textregistered\, image processing toolbox. The measured contact angle is about$30^\circ$. Photographs are taken after the filament has broken up into droplets and the resulting droplets reached an equilibrium state.}
\caption{Breakup and no-breakup results plotted as a function of Oh versus AR: breakup ({\textcolor{red}{$\bullet$}}) and no-breakup ({\textcolor{green}{$\circ$}}). Experimental results in \cite{pita}: breakup ({\textcolor{red}{$\times$}}) and no-breakup ({\textcolor{green}{$\times$}}).}
\caption{Breakup and no-breakup results plotted as a function of Oh versus AR: breakup ({\textcolor{red}{$\bullet$}}) and no-breakup ({\textcolor{green}{$\circ$}}), for $\lambda=1$ (a), $0.1$ (b), $0.01$ (c), and $0$ (d).}
\caption{\textbf{Spin-texture of metallic surface states.} \textbf{a${-}$d}, Spin polarization recorded along the $\overline{\Gamma}$ states showing the helical spin texture, for $t\,{=}\,2.5$ nm. The numbers are the net spin polarization values. %and signs are given with respect to $k_{x}$ and $k_{y}$ directions. \textbf{e}, Band structure along $\overline{\Gamma}{-}\overline{K}$ for $t\,{=}\,5$ nm. Red solid lines are $\overline{\Gamma}$ ($k_{[\bar{1}10]}\,{\approx}\,0.1$ $\textup{\AA}^{-1}$) and $\overline{M}'$ ($k_{[\bar{1}10]}\,{\approx}\,0.7$ $\textup{\AA}^{-1}$) hole states of Bi(110) respectively. Dotted lines crossing the Fermi level correspond to the electron pockets near the $\overline{X_{1}}'$ point shown in Figs.~\ref{Fig2}\textcolor{blue}{$\,$f,$\,$g}. \textbf{f,$\,$g}, Spin polarization of $\overline{M}'$ states.}
\caption{\label{fig:independenceF} (a) Levels of independence $p$ as functions of the noise strength $T$ for different fixed values of the utilities \textcolor{mycolor}{$(u^I;u^C)$}. (b) Time evolution of the average levels of independence in the system of the size $N=10^6$. The initial utilities are as follows: \textcolor{mycolor}{$u^I_0=0.4$ and $u^C_0=0$}. Markers represent the outcome of Monte Carlo simulations and are connected just to guide the eye. Each line refers to a different value of $T$. The noise increases from the top to the bottom. Note that the final values of $\langle p\rangle$ depend on the noise level, in contrast to the case with the same initial utilities where $\langle p\rangle=0.5$ independently of the noise.}
\caption{\label{fig:histogramSym} Normalized histograms of the independence level $p$ in the system comprised of $N=10^6$ agents after $10^3$ MCS for the same initial utilities \textcolor{mycolor}{$u_0^I=u_0^C=0$} at five different noise levels increasing from the left to the right. The group of influence contains $q=4$ agents. Note that although the distribution changes with $T$, it stays symmetric all the time. Moreover, the average value of independence in the system $\langle p\rangle$ is not affected by the noise. It remains unchanged for different $T$ and equals $\langle p\rangle=0.5$.}
\caption{\label{fig:histogramPanel} Normalized histograms of the independence level $p$ in the system comprised of $N=10^6$ agents with the bias in the initial values of utilities at five different noise levels increasing from the first column on the left to the last column on the right. The top row corresponds to the following initial values \textcolor{mycolor}{$u^I_0=0.4$ and $u^C_0=0$} while the bottom row refers to the values \textcolor{mycolor}{$u^I_0=0$ and $u^C_0=0.4$}. The data are collected after $10^3$ Monte Carlo steps.}
\caption{\label{fig:phaseDiagramsQVoter} Phase diagrams for the $q$-voter model in the (a) quenched and (b) annealed approaches. Gray lines indicate the stationary values of the up-spin concentration given by Eqs.~(\ref{eq:qurnchedc}) and (\ref{eq:annealedc}), continuous and dashed ones refer to stable and unstable states accordingly. (a) For the quenched case, when the initial utilities are equal, we have $\bar{p}=0.5$ (dotted vertical line) for all $T$. Lower values of the average level of independence are related to the situation in which \textcolor{mycolor}{$u_0^I<u_0^C$}. Higher values, on the other hand, correspond to the following utilities \textcolor{mycolor}{$u_0^I>u_0^C$}. For these assymetric initial conditions, arrows illustrate how $\bar{p}$ will change when we increase the noise $T$ in the system. (b) For the annealed case, the average level of independence in the system amounts to $\bar{p}=0.5$ (dotted vertical line), and the only stable value of the stationary up-spin concentration is $c=0.5$. Note that for the annealed case, $\bar{p}$ corresponds to the probability of an agent being independent in a given time step whereas for the quenched one, $\bar{p}$ refers to the fraction of agents in the system that are independent all the time.}
\caption{\label{fig:phaseDiagrams} Phase diagrams for the $q$-voter model with extreme value memory and exponentially distributed utilities with parameter $\lambda=1$. \textcolor{mycolor}{Initially, all spins are up, so $c(0)=1$.} Dots represent the outcome of Monte Carlo simulations of the system containing $N=10^4$ agents. The results are averaged over $10^2$ runs and collected after $10^3$ MCS. Dotted lines are drawn just to guide the eye. (a) The model with no bias in the initial values of the utilities, i.e., \textcolor{mycolor}{$u_0^I=u_0^C=0$} for different $q$-panel sizes. (b) The model with the group of influence comprised of $q=4$ agents and different initial conditions for the utilities \textcolor{mycolor}{$(u_0^I; u_0^C)$}.}
\caption{\red{[These results are directly tested on Test Data.]} \blue{[Parenthesized are test data acc on best dev model]} Performance of the entity-based coherence models in \textbf{Discrimination} and \textbf{Insertion} tasks on \textbf{CNET} forum conversations for \textbf{Temporal}, \textbf{Path-level} and \textbf{Tree-level} representations. Superscript $\dagger$ indicates a \textbf{neural} model is significantly superior to its \textbf{non-neural} counterpart with p-value $<0.01$, and $\bigstar$ indicates a \textbf{path or tree} level model is significantly superior to its \textbf{temporal} counterpart with p-value $<0.01$,}
\caption{Scaling of the Fisher information: (a) effective Fisher information for the distinguishability parameter $\epsilon$. The points correspond to numerical results, and the solid lines correspond to $2N^2$ (black) and $N$ (red). (b) effective Fisher information for the phase $\phi$. The points correspond to numerical results, and the solid lines correspond to $2N(N+1)$ (black) and $N$ (red). (c) Trade-off in the optimality of individual estimations quantified by $\Upsilon$. In all plots: {\color{blu} $\bullet$}: $\epsilon=0.14$, {\color{ocra} $\blacksquare$}: $\epsilon=0.23$, {\color{verde} $\blacklozenge$}: $\epsilon=0.32$, {\color{rosso} $\blacktriangle$}: $\epsilon=0.50$, {\color{viola} $\blacktriangledown$}: $\epsilon=1$.}
\caption{A street measurement map in NJ suburb. \textcolor{\editcolor}{The outdoor CPE (indicated by the red circle) was aimed 45$^\circ$ to illuminate the street, and the indoor CPE (blue triangle) was aimed perpendicular to the window}. The rotating horn ``base station'' moved along the street at ranges from 20 to 200 meters (orange line trajectory).}
\caption{CDFs of measured building penetration loss for three representative single-family homes. The median was found to be 8.9 dB, 15.1 dB and 17.1 dB, respectively, as compared to the 18.6 dB median of the 3GPP low loss model. The difference between the median and mean values is within 0.2 dB, in line with the 0.15 dB power resolution of the sounder. \textcolor{\editcolor}{In all three cases the measured samples are well represented by log-normal distributions (thin solid lines) with standard deviation of 3.0, 2.5, and 2.8 dB, respectively.} }
\caption{(Color online) Phase space trajectories governed by the effective reduced ODE model (given by\textcolor{blue}{{} }Eqs. (\ref{vp_model})) providing a variational description for the direct reflection type of solitons scattering. In the three cases considered ($\gamma=0.5$, $1.0$, $1.5$), a total of $8$ trajectories with $p_{0}=10$ and $v_{0}'\in(0,v_{c}^{\text{\tiny R}}(\gamma)]$ are plotted. The background is a contour line plot of the function $\partial_{p}G_{0}'.$ Since $\Gamma=-0.04<0$, the attraction zones are the highlighted by gray/black regions, corresponding to $\text{sgn}\left(\partial_{p}G_{0}'\right)=-1$, while repulsion zones are identified by the white regions, corresponding to $\text{sgn}\left(\partial_{p}G_{0}'\right)=+1$.}
\caption{ \textbf{Experimental darkfield spectra training data-set for 4 bits.} Our data comprises measurements from 625 copies for each of the 16 ``4 bit'' geometries (this makes a total of \(625\times 16\times 2 = 20000\) acquired spectra). The spectra are superposed above each other. Blue lines: DF scattering for \(X\)-polarized light, orange lines: \(Y\)-polarization. Insets show SEM images of one representative copy of the respective nanostructure, the areas are \(600 \times 600\,\)nm\(^2\) large, the scalebar in ``1111'' is \(200\,\)nm. The not shown 2 bit dataset consists of 5000, the case of 3 bits comprises 10000 and the 5 bit set contains 40000 scattering spectra (all shown together with SEM images in the {\color{blue}supporting informations}, (Figs.~S4-S11\comment{CHECK}), where also a separate 4 bit dataset and simulations are shown in order to demonstrate the reproducibility (Figs.~S12-S14\comment{CHECK}). }
\caption{ \textbf{Accuracy of network trained on reduced spectral information.} (a) Training using scattering from spectral window of reduced width for the 2, 3, 4 and 5 bit datasets (left top to right bottom plot). The scattering intensity was taken from a window either at the short wavelength side (blue lines), in the center (green lines) or from the red edge of the spectra. (b) Average (top plot) and worst-digit (bottom pot) read-out accuracy using a network trained on the scattering intensity of \(X\)- and \(Y\)-polarized light at a discrete number of wavelengths. The explicit positions of \(\lambda_i\) for the different cases are given in the Methods section. t-SNE plots for all reduced spectral information datasets can be found in the {\color{blue}SI}, Figs.~S15.\comment{CHECK} }
\caption{ \textbf{Neural-network based data read-out via the RGB color values.} (a) polarization filtered dark-field color images of representative \(3\times 3\) arrays of the ``4 bit'' digit structures. Left: \(X\)-polarization, right: \(Y\)-polarization. The DF images show areas of \(7.5\times 7.5\,\)\textmu m\(^2\). The insets show the average RGB color of the \(3\times 3\) structures. (b) scheme of the fully connected artificial neural network used for the RGB classification task. (c) t-SNE\cite{van_der_maaten_visualizing_2008} visualization of the ``3 bit'' training sets for only \(X\), only \(Y\), \(X\) plus \(Y\) or \(X\), \(Y\) and scattered intensity \(I\) (from left to right). Only if using simultaneously both polarizations (``XY'' and ``XY+I''), the different bit sequences show a clear separation in the t-SNE plots. (d) t-SNE plots for the XY+I cases of the 2, 3, 4 and 5 bit training sets (from left to right). (e) information retrieval accuracy of the network, trained on the different data-sets consisting of only the \(X\)-filtered, only the \(Y\)-filtered, \(X\)+\(Y\) filtered and \(XY\) + the scattered intensities \(I\). t-SNE plots for all data-sets can be found in the {\color{blue}SI}, Fig.~S15.\comment{CHECK} }
\caption{ \textbf{9 bit per nanostructure information encoding.} (a) geometry of a silicon nanostructure encoding \(3\times 3 = 9\) bits (512 possible combinations). Each silicon block occupies an area of \(105\times 105\,\)nm\(^2\). The L-shaped sidewall is \(45\,\)nm wide. The height is \(90\,\)nm. An SEM image of a fabricated structure is given in the inset, where the scalebar is 200\,nm. The shown example represents the decimal number ``321''.% (b) selected examples, illustrating the training data generation by a numerical expansion of the experimental spectra. SEM images show areas of \(550\times 550\,\)nm\(^2\). We fabricated 4 copies of each possible 9 bit nanostructure. Via random superposition of the experimental scattering spectra of these 4 copies, we generate a large set of spectra, allowing us to train and test the performance of the binary information readout ANN (see also {\color{blue}SI} Fig.~S23\comment{CHECK} for more details). % (c) accuracy of the ANN trained on the experimental 9 bit data using the full spectra (left, the solid line is a guide to the eye) or scattering intensities at a limited number of discrete wavelengths (right). In both cases, \(X\) and \(Y\) polarized data is used simultaneously. % (d) evaluation of the robustness of 9 bit read-out with respect to noise on fully numerical data. We use simulated scattering spectra for all 9 bit geometries with different amounts of random noise (see {\color{blue}SI} Fig.~S33\comment{CHECK} for details on the training data generation). Example spectra for noise levels from (i) 10\% to (iv) 25\% are shown at the example of structure ``001011010'' (decimal 90). The plot on the left shows the readout error rate on the numerical data as function of noise level (training on the full spectra of both polarizations). % (e) scattering deviation relative to the isolated structure as function of the distance between two digit encoding structures for focused illumination. For the estimation of the feasible information density, 10\% deviation from the unperturbed spectrum are assumed to be tolerable (red dashed horizontal line). This leads to an information density (green indicator) around \(40\)\% higher than the blue-ray disc (blue indicator). The 4 bit structures yield about \(75\)\% of the blue-ray density. }
\caption{Local density of states of 3D LiF, obtained within second order perturbation theory, using the XXX basis. Blue traces: Analytically continued local spectral function of the near-gap states. Black traces: Analytical continuation of the $k$-space spectral function, followed by a $k$-summation to obtain the local part. \textcolor{red}{Left label font, left axis font bigger. Same with bottom axis. Bottom axis should be $\omega [eV]$. Legend should be: $A_\text{loc}(\omega)$ from $A_k(\omega)$. Then $A_\text{loc}(\omega)$ from $G_\text{loc}(i\omega_n).$ We will explain rest in the text.} }
\caption{Examples of generated summaries on the Gigaword corpus. \textbf{D}: source document, \textbf{R}: reference summary, \textbf{OR}: output of the \textbf{R}einforced-ConvS2S model, \textbf{OT}: output of the Reinforced-\textbf{T}opic-ConvS2S model. The words marked in {\color{blue}blue} are topic words not in the reference summaries. The words marked in {\color{red}red} are topic words neither in the reference summaries nor in the source documents.}
\caption{Examples of generated summaries on the LCSTS dataset. \textbf{D}: source document, \textbf{R}: reference summary, \textbf{OR}: output of the \textbf{R}einforced-ConvS2S model, \textbf{OT}: output of the Reinforced-\textbf{T}opic-ConvS2S model. The words marked in {\color{blue}blue} are topic words not in the reference summaries. The words marked in {\color{red}red} are topic words neither in the reference summaries nor in the source documents. All the texts are carefully \textit{translated from Chinese}. }
\caption{Examples of generated summaries on the LCSTS dataset. \textbf{D}: source document, \textbf{R}: reference summary, \textbf{OR}: output of the \textbf{R}einforced-ConvS2S model, \textbf{OT}: output of the Reinforced-\textbf{T}opic-ConvS2S model. The words marked in {\color{blue}blue} are topic words not in the reference summaries. The words marked in {\color{red}red} are topic words neither in the reference summaries nor in the source documents. All the texts are carefully \textit{translated from Chinese}.}
\caption{The complex $\CG$ where $\G$ is the graph consisting of a single vertex. The \textcolor{blue}{fundamental domain} of the action of $\AG$ on $\CG$ is the edge spanned by $A_{\{s\}}$ and $A_\emptyset$. }
\caption{The fundamental domain of the complex $\CG$ where $\G$ is the graph shown, with \textcolor{blue}{the link of $A_\emptyset$.} By Lemma \ref{link iso to Gamma} the flag complex $\tilde{\G}$ is isomorphic to $lk(A_\emptyset)$. }
\caption{Performance improved by background-aware block-wise random synthesis. $\blacksquare$ indicates CTPN, {\color{blue}$\star$} indicates applying the predicted guidance mask on CTPN during testing without retraining, {\color{green}$\blacklozenge$} indicates Guided CTPN trained with the predicted guidance mask, {\color{magenta}{$\blacktriangledown$}} indicates Guided CTPN trained with the predicted guidance mask and random synthesis, and {\color{red}{$\large\bullet$}} indicates Guided CTPN trained with the ground truth mask and random synthesis.}
\caption{Qualitative results of the proposed Reciprocal Attention Fusion mechanism for Visual Question Answering. Given a question and an image (columns: $1,4$), attention based on image-grid (columns: $2,5$) and object proposals (columns: $3,6$) is shown above. Correct and incorrect answers are shown in \textcolor{green}{green} and \textcolor{red}{red}, respectively. Remarkably, the two attention levels provide complementary information about localized regions and objects that in turn help in obtaining the correct answer (rows: $1,2,3,4$). In some failure cases of our technique, ambiguous attention maps lead to incorrect predictions (row: $5$).}
\caption{{\bf Phase diagram of the $r$-GRN model.} Two transition points $p_{b}$ \textcolor{blue}{$\triangle$} and $p_{c}$ \textcolor{red}{$\bigcirc$} are determined for various $g$. $n_s(p,g)$ decays following a power law with $\tau > 3$ in the infinite-order critical region and $2 < \tau < 3$ in the second-order critical region. Thus, the mean cluster size is finite and diverges in those regions, respectively. As $g$ approaches one, the two transition points are closer and converge to the critical point of an infinite-order transition, represented by \textcolor{orange}{$\blacksquare$}.}
\caption{\small \textbf{Analysis of gender bias:} Box plot of the score differences on the gender sentence pairs for each system on the valence regression task. Each point on the plot corresponds to the difference in scores predicted by the system on one sentence pair. {\color{blue} \ding{115}} represents F$\uparrow$--M$\downarrow$ significant group, {\color{orange} \ding{116}} represents F$\downarrow$--M$\uparrow$ significant group, and {\color{green} \ding{108}} represents F=M not significant group. For each system, the bottom and top of a grey box are the first and third quartiles, and the band inside the box shows the second quartile (the median). The whiskers extend to 1.5 times the interquartile range (IQR = Q3 –- Q1) from the edge of the box. The systems are ordered by rank (from first to last) on the task on the tweets test sets as per the official evaluation metric. }
\caption{\small \textbf{Analysis of race bias:} Box plot of the score differences on the race sentence pairs for each system on the valence regression task. Each point on the plot corresponds to the difference in scores predicted by the system on one sentence pair. {\color{blue} \ding{115}} represents AA$\uparrow$--EA$\downarrow$ significant group, {\color{orange} \ding{116}} represents AA$\downarrow$--EA$\uparrow$ significant group, and {\color{green} \ding{108}} represents AA=EA not significant group. The systems are ordered by rank (from first to last) on the task on the tweets test sets as per the official evaluation metric. }
\caption{\small \textbf{Analysis of gender bias:} Box plots of the score differences on the gender sentence pairs for each system on the four emotion intensity regression tasks. Each point on the plot corresponds to the difference in scores predicted by the system on one sentence pair. {\color{blue} \ding{115}} represents F$\uparrow$--M$\downarrow$ significant group, {\color{orange} \ding{116}} represents F$\downarrow$--M$\uparrow$ significant group, and {\color{green} \ding{108}} represents F=M not significant group. The systems are ordered by their performance rank (from first to last) on the task as per the official evaluation metric on the tweets test sets. The system with the lowest performance had the score differences covering the full range from -1 to 1, and is not included in these plots. }
\caption{\small \textbf{Analysis of race bias:} Box plots of the score differences on the race sentence pairs for each system on the four emotion intensity regression tasks. Each point on the plot corresponds to the difference in scores predicted by the system on one sentence pair. {\color{blue} \ding{115}} represents AA$\uparrow$--EA$\downarrow$ significant group, {\color{orange} \ding{116}} represents AA$\downarrow$--EA$\uparrow$ significant group, and {\color{green} \ding{108}} represents AA=EA not significant group. The systems are ordered by their performance rank (from first to last) on the task as per the official evaluation metric on the tweets test sets. The system with the lowest performance had the score differences covering a much larger range (from -0.3 to 0.3), and is not included in these plots. }
\caption{Detailed IoU of ContextNet (cn14) compared to version with half (cn12) and eighth (cn18) resolution, and its multi-level version (cn124). Small-size classes (\textcolor[rgb]{0,0.7,0}{green}), classes with fine detail (\textcolor[rgb]{0,0,1}{blue}), and classes with very few samples (\textcolor[rgb]{1,0,0}{red}) benefit from high resolution.}
\caption{{\it Magnetosonic}. Real (upper) and imaginary (lower) parts of the magnetosonic dispersion relation, shown for different electron betas $\beta_e$. We set $\theta=60^\circ$ and $C_0=0.234$ for all calculations, but vary the electron beta $\beta_e$. The distribution is unstable in the range $0.3 \lesssim \beta_e \lesssim 0.7$, \red{with monotonically increasing growth as $\beta_e$ increases in this range.}}
\caption{\red{ {\it KAW---Less oblique}. Dispersion relations for the original core-strahl model (solid lines, from Fig.~\ref{theta_scan_kaw_plot}) and the core-halo-strahl model (dashed lines) described in this appendix.}}
\caption{\red{ {\it KAW---More oblique}. Dispersion relations for the original core-strahl model (solid lines, from Fig.~\ref{theta_scan_kaw_oblique_plot}) and the core-halo-strahl model (dashed lines) described in this appendix.}}
\caption{ \red{{\it Magnetosonic}. Dispersion relations for the original core-strahl model (solid lines, from Fig.~\ref{theta_scan_ms_plot}) and the core-halo-strahl model (dashed lines) described in this appendix.}}
\caption{\red{ {\it Magnetosonic}. Dispersion relations for the original core-strahl model (solid lines, from Fig.~\ref{beta_scan_ms_plot}) and the core-halo-strahl model (dashed lines) described in this appendix.}}
\caption{Left-ventricle segmentation results with different levels of supervision. \rred{Bold font highlights the best weakly supervised setting.}}
\caption{\rred{Evolution of the DSC during training for the left-ventricle validation set, including the weakly supervised learning models and different strategies analyzed, with also the full-supervision setting. As tags and common bounds achieve similar results, we plot only common bounds for better readability.}}
\caption{\rred{Ablation study on the amounts of fully and weakly labeled data. We report the mean DSC of all the testing cases, for all the settings and using the same architecture.}}
\caption{\rred{Mean DSC values over the number of fully annotated patients employed for training.}}
\caption{\bblue{Mean Dice scores (DSC) for several degrees of supervision}\rred{, using the vertebral-body and prostate validation sets. Bold font indicates the best weakly supervised setting for each data set.}}
\caption{\rred{Qualitative comparison of the different methods using examples from the LV dataset. Each column depicts segmentations obtained by different methods, whereas each row represents a 2D slice from different scans (Best viewed in colors).}}
\caption{\rred{Qualitative comparison using examples from the VB dataset. Each column depicts segmentations obtained by different levels of supervision, whereas each row represents a 2D slice from different scans (Best viewed in colors).}}
\caption{\bblue{Qualitative comparison of the different levels of supervision. Each row represents a 2D slice from different scans. (Best viewed in colors)}}
\caption{\bblue{Ablation study on the lower and upper bounds of the size constraint} \rred{using the vertebral body dataset.}}
\caption{Training times for the diverse supervised learning strategies \rred{with a batch size of 1}, using tags and size constraints.}
\caption{We first generate the meaning sketch~$a$~for natural language input~$x$. Then, a fine meaning decoder fills in the missing details (shown in {\color{red} red}) of meaning representation~$y$. The coarse structure~$a$~is used to guide and constrain the output decoding. }
\caption{\label{fig:Devicelayout}\textbf{ Two-qubit device layout and operation.} \textbf{a}, False colour scanning electron microscope image of the device. Two quantum dots D1 and D2 are formed underneath gates G1 (blue) and G2 (red). The gates CB (purple), G3 and G4 (grey) form confinement barriers that laterally define the quantum dots. RG (yellow) is the reservoir gate that supplies electrons to the quantum dots. The gate electrodes ST, SLB and SRB (green) define a single electron transistor, designed to sense charge movement in the quantum dot region. An AC current running through the ESR line (light blue) generates an oscillating magnetic field to manipulate the electron spins. The direction of the external magnetic field $B_{0}$ is indicated by the white arrow. \textbf{b-e}, Control path in the charge stability diagram and schematic depicting initialization and readout: \textit{(i.)} Load a spin-down electron from the reservoir into D2 by biasing to the (0,0)-(0,1) transition (I1) for 2.75~ms. \textit{(ii.)} Move the electron to a spin relaxation hot-spot~\cite{Yang2013} (H) close to the (0,1)-(1,0) anti-crossing and keep it there for $300$~$\mu$s to improve the initialization fidelity~\cite{Watson2018,Zhao2017}. Then transfer the electron to D1 by moving it through the anti-crossing~\cite{Baart2016}, completing the initialization of qubit \textcolor[rgb]{0,0,1}{Q1} as $\ket{\downarrow}$ (I1-H-I2 in the stability diagram). \textit{(iii.)} Load another spin-down electron into D2 by biasing to the (1,0)-(1,1) transition (I2) for 2.75~ms to initialize qubit \textcolor[rgb]{1,0,0}{Q2} as $\ket{\downarrow}$. This sequence initializes the system $\ket{\textcolor[rgb]{0,0,1}{Q1},\textcolor[rgb]{1,0,0}{Q2}}$ as $\ket{\downarrow\downarrow}$. The two-qubit system is now ready for operation. \textit{(iv.)} Perform single-qubit and two-qubit quantum operations on the two qubits in the (1,1) region (C) by using sequences of selective ESR pulses. \textit{(v.)} Read out the qubit \textcolor[rgb]{1,0,0}{Q2} at the (1,0)-(1,1) transition via spin-dependent tunneling~\cite{Elzerman2004} by biasing to R2 for 2.75~ms, then ensure D2 is unloaded by pulsing deeper into the (1,0) region (U) for 3~ms. \textit{(vi.)} Transfer the qubit \textcolor[rgb]{0,0,1}{Q1} from dot D1 to D2 by adiabatically sweeping through the (1,0)-(0,1) anti-crossing within 5~$\mu$s, which is fast enough to avoid relaxation at the hot spot~\cite{Baart2016}. \textit{(vii.)} Read out \textcolor[rgb]{0,0,1}{Q1} at the (0,0)-(0,1) transition (R1) for 2.75~ms. This concludes the operational sequence. In our devices, performing readout by shuttling \textcolor[rgb]{0,0,1}{Q1} from D1 to D2 is advantageous over reading out \textcolor[rgb]{0,0,1}{Q1} at the (1,0)-(0,0) transition directly, due to the slow tunneling rate from D1 to the reservoir.}
\caption{\label{fig:Rabi}\textbf{ Independent and Conditional Two-Qubit Control.} \textbf{a} ESR spectra of the two-qubit system. Here, the peaks are power-broadened and the linewidths are given by the respective Rabi frequencies. We prepare $\ket{\textcolor[rgb]{0,0,1}{Q1},\textcolor[rgb]{1,0,0}{Q2}}$ in either $\ket{\downarrow\downarrow}$, $\ket{\tilde{\downarrow\uparrow}}$ or $\ket{\tilde{\uparrow\downarrow}}$ and measure the spin-up probability of \textcolor[rgb]{0,0,1}{Q1} and \textcolor[rgb]{1,0,0}{Q2} as a function of the applied microwave frequency. Four distinct ESR peaks arise due to the presence of a finite exchange coupling $J$ and a Zeeman energy difference $\delta E_\textrm{Z}$, centered around $f_c=\bar{E_Z}/h=39.33$~GHz. Each resonance peak represents a rotation of the target qubit conditional on the state of the control qubit (CROT signal). The energy level diagram (inset) maps each peak to the corresponding transition between a pair of two-qubit eigenstates. \textbf{b} Controlled qubit rotations are naturally implemented by pulsing at individual resonance frequencies. A first pulse $U^{\tau}_{2\downarrow}$ performs Rabi rotations on \textcolor[rgb]{1,0,0}{Q2} that result in the resonance frequency of \textcolor[rgb]{0,0,1}{Q1} oscillating between $f_{1\downarrow}$ and $f_{1\uparrow}$. \textbf{c} Independent qubit control can be achieved under the presence of constant $J$ by applying microwave pulses at the two conditional frequencies simultaneously. A first pulse at $U_{1\uparrow}^{\tau_1}U_{1\downarrow}^{\tau_1}$ defines the state \textcolor[rgb]{0,0,1}{Q1}. A second pulse $U_{2\uparrow}^{\tau_2}U_{2\downarrow}^{\tau_2}$ then rotates \textcolor[rgb]{1,0,0}{Q2}, independent of the state of \textcolor[rgb]{0,0,1}{Q1}. }
\caption{\label{fig:RBM}\textbf{ Two-qubit randomized benchmarking.} \textbf{a} Primitive gates $X/2$, $X/2+$CROT, Z-CROT and CROT acting on \textcolor[rgb]{1,0,0}{Q2} and their corresponding ESR pulse sequences. Together with the virtual gate $Z_V/2$ and the gates acting on \textcolor[rgb]{0,0,1}{Q1}, these gates span the two-qubit Clifford space. \textbf{b} Projected state probability as a function of the number of Clifford gates in each sequence. Each sequence is repeated 125 times and the measurement averages over 51 sequences of the same length. The Clifford gates are randomly chosen from 11520 elements of the two-qubit Clifford group, with the $L$-th gate projecting the state to the $\ket{\uparrow\uparrow}$ state. The extracted Clifford fidelity is $F_{\mathrm{Clifford}}=94.7\pm0.8$~\%, the primitive gate fidelity is $F_{\mathrm{primitive}}=98.0\pm0.3$~\%, and the conditional $\frac{\pi}{2}$-pulse fidelity is $F_{\frac{\pi}{2}}^{\mathrm{cond}}=99.0\pm0.15$~\%. }
\caption{The architecture of Cycle-Dehaze Network where $G$ \&$F$ refers to the generators, and $D_x$ \&$D_y$ to the discriminators. For the sake of clarity, the representation is split into two parts: hazy to clean image, and clean to hazy image. {\color{red} Best view in color.} }
\caption{Average PSNR and SSIM results on NYU-Depth~\cite{NYUdataset} dataset. Most of the accuracies taken from the paper~\cite{yang2018towards}. Numbers in {\color{red}{red}} and {\color{blue}{blue}} indicate first and second best results, respectively. The second column of the table shows the values which are average PSNR and SSIM results calculated directly between the each hazy and its ground truth image.}
\caption{ Log-odds ratios of politeness strategies and prompt types exhibited in the first and second comments of conversations that turn awry, versus those that stay on-track. \textbf{All}: {\textcolor{darkorchid}{Purple}} and {\textcolor{seagreen}{green}} markers denote log-odds ratios in the first and second comments, respectively; points are solid if they reflect significant (\mbox{$p < 0.05$}) log-odds ratios with an effect size of at least 0.2. % \textbf{A}: \textcolor{darkorchid}{$\diamondsuit$}s and \textcolor{seagreen}{$\square$}s denote \textbf{first} and \textbf{second} comment log-odds ratios, respectively; * denotes statistically significant differences at the \mbox{$p < 0.05$} (*), \mbox{$p < 0.01$} (**) and \mbox{$p < 0.001$} (***) levels for the first comment (two-tailed binomial test); + denotes corresponding statistical significance for the second comment. \textbf{B} and \textbf{C}: $\triangledown$s and $\bigcirc$s correspond to effect sizes in the comments authored by the \textbf{attacker} and \textbf{non-attacker}, respectively, in \textbf{attacker initiated} (\textbf{B}) and \textbf{non-attacker initiated} (\textbf{C}) conversations. }
\caption{The precisions of the top 100 retrieved documents for NASH-DN with \textcolor{blue}{\emph{stochastic}} or \textcolor{red}{\emph{deterministic}} binary latent variables. }
\caption{From top to bottom: panels 1 to 6: UVOT data ($w1$, $m2$, $w2$, $u$, $b$, $v$ bands observed flux in units of mJy) acquired during the observations described in Table~\ref{xrtdatalog}; panel 7: {\it Swift}/XRT ($0.3-10.0$\,keV observed flux in units of$10^{-12}$\,erg\,cm$^{-2}$\,s$^{-1}$); last panel: \gray 48h-bin light curve ($> 100$\,MeV observed flux in units of$10^{-7}$\,photons\,cm$^{-2}$\,s$^{-1}$), AGILE data in red, \lat in black. The only NIR data from OAGH available during this time interval were taken on MJD $\sim$ 57498 (flux density = $(2.64 \pm 0.10)$ mJy in J band, $(4.16 \pm 0.19)$ mJy in H band, $(5.91 \pm 0.27)$ mJy in Ks band).}
\caption{Differential spectrum for the second \gray flare, ten-day integration, from 2016 April 10 UT 12:00:00 to 2016 April 20 UT 12:00:00 (MJD: 57488.50 -- 57498.50). Blue points: \grid data, 50 MeV -- 3 GeV. Red points: \lat data, 100 MeV -- 300 GeV. }
\caption{Imaging of low-conductivity samples with an RF atomic magnetometer operating at \SI{2}{\MHz}. $\Delta \Phi$ scans of the doped Si specimens. {\bf (a)} $\sigma=\SI{1e4}{\siemens\per\metre}$. {\bf (b)} $\sigma=\SI{5e3}{\siemens\per\metre}$. {\bf (c)} $\sigma=\SI{1e3}{\siemens\per\metre}$. {\bf (d)} $\sigma=\SI{5e2}{\siemens\per\metre}$. {\bf (e)} $\sigma=\SI{1e2}{\siemens\per\metre}$. {\bf (f)} $\sigma=\SI{5e1}{\siemens\per\metre}$. \red{ {\bf (g)} Simultaneous imaging at \SI{2}{\mega\hertz} of specimens with conductivities spanning $[\SI{1e4}{\siemens\per\metre}, \SI{5e1}{\siemens\per\metre}]$. Conductivity decreases from top left to right bottom. Samples are not in electrical contact.}}
\caption{Near-resonant imaging. {\bf (a)} Magnetometer amplitude response near \SI{2}{\mega\hertz}. The non-linear Zeeman effect causes the RF resonance to split in four components, \red{corresponding to individual transitions $|F=2, m_{F}\rangle \leftrightarrow |F=2, m_{F}\pm1\rangle$, as indicated by the respective labels.} The shaded area highlights the region explored for near-resonant operation. {\bf (b)} Magnetometer phase profile and operating points for near-resonant imaging. Inset: the entire phase response highlighting the near-resonant region. {\bf (c)} Example of the measured phase noise ($\delta\Phi_{\text{meas}}$) across the imaging area \new{(the total phase noise in an image without specimens)} versus detuning. {\bf (d)} Calculated phase noise ($\delta\Phi_{5\text{nT}}$) produced by \SI{5}{\nano\tesla} magnetic field noise, as measured across the imaging area. \red{The circled numbers mark the position with respect to the resonance curve in panel (b).} The vertical dotted lines mark the configuration used throughout this work.}
\caption{\red{Imaging tunability. \textbf{(a)-(c)} Images of the \SI{5e2}{\siemens\per\metre} sample at \SI{200}{\kilo\hertz}, \SI{800}{\kilo\hertz}, and \SI{1.4}{\mega\hertz}. \textbf{(d)} Maximum phase change across an image ($\Delta \Phi_{\text{max}}$) for each of the six doped Si samples as a function of frequency. Dashed lines are the best interpolating curves (piecewise cubic Hermite interpolating polynomials) for each dataset.}}
\caption{\powerpredictor system: Features used\label{table:cb_features}}
\caption{Sample generations from an RNN language model (LM) and our system (L2W) conditioning on the context shown on the top. % on the BookCorpus dataset. The {\color{myred}\underline{red, underlined}} text highlights %egregious repetitions, while the \textit{\color{myblue}blue, italicized} text highlights details that have a direct semantic parallel in the reference text. }
\caption{Proposed model architectures, showing a single training instance $(\mathbf{x}_i, y_i)$ with domain $d_i$, and baseline model with domain adversarial loss. $\operatorname{CNN}$ denotes a convolutional network, $\operatorname{D}$ indicates a discriminator ($d$ for domain adversarial and $g$ for domain generation), and \textcolor{red!60!white}{red dashed} and \textcolor{blue}{blue} lines denote adversarial and standard loss, resp.}
\caption{\small Recovery time analysis of Transaction Logger at varying fault timing. \cred{Can you split this figure into two figures as shown in Fig4 and 5 for each workload? This figure has two (a)s, (b)s, etc. } }
\caption{\small Recovery time analysis of Universal Logger at varying fault timing. \cred{In (d), y-axis scale is different from that of (a-c).} \cred{Separate this one figure into two figures, one for big files, and the other for small files as for Fig 4 and 5. There are two (a-d)s. } \label{fig:bbcpvsUL} \vspace{0.1in} \end{figure*} \end{comment} Minimizing the recovery time upon resuming from fault is one of the major objectives in our FT-LADS design. In this section, we have evaluated the FT-LADS recovery time for small and big workloads and compared it against bbcp data transfer tool. For effective evaluation of recovery time of proposed fault tolerance methods, we created a simulation environment in which we generate faults after transferring 20\%, 40\%, 60\%, 80\% of total data size. As faults can occur at any end of the transfer, we can simulate the faults at either source or sink. However, for the purpose of our experiments, we have executed this simulation in the source end. %\sout{Using the environment described in Section~\ref{sec:expr} for measuring recovery time for both bbcp and FT-LADS. \cred{This sentence is not complete. Please check and rephrase it.}} %{\color{blue} Using the experimental environment, described in Section~\ref{sec:expr}, we measured the recovery time for both bbcp and FT-LADS. %} On these hosts, LADS uses CCI's Verbs transport, which natively uses the underlying InfiniBand interconnect. Whereas, bbcp uses the IPoIB interface which supports traditional sockets. In LADS, varying the number of I/O threads maximizes CPU utilization on the data transfer node. However, bbcp uses configurable window size and multiple streams to improve the performance. Based on the experimental results presented in LADS \cite{kim:fast15}~\cite{lads17}, LADS data transfer performance increases linearly with the number of I/O threads. Whereas, bbcp has less impact while increasing the number of tcp streams. For fair performance comparison between the two, we have configured FT-LADS to use $4$ I/O threads and bbcp to use $2$ tcp streams with window size of 8MB. Our experiments are designed to calculate the transfer time before and after fault. Based on these times and using Equation~\ref{equ1}, we estimated the recovery time. \begin{figure*}[!t] \centering \begin{tabular}{@{}cccc@{}c@{}c@{}c@{}c@{}} \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_20_1GB.pdf}& \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_40_1GB.pdf}&\\ \small (a) Big loads 20\% fault time & \small (b) Big loads 40\% fault time &\\ \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_60_1GB.pdf}& \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_80_1GB.pdf}\\ \small (c) Big loads 60\% fault time & \small (d) Big loads 80\% fault time \\ \end{tabular} % \vspace{0.1in} \caption{\small Recovery time analysis of FileLogger at varying fault timing for big workloads. } \label{fig:bbcpvsFL_big} % \vspace{0.1in} \end{figure*} \begin{figure*}[!t] \centering \begin{tabular}{@{}cccc@{}c@{}c@{}c@{}c@{}} \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_20_1MB.pdf}& \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_40_1MB.pdf}&\\ \small (a) Small loads 20\% fault time & \small (b) Small loads 40\% fault time &\\ \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_60_1MB.pdf}& \includegraphics[width=0.35\textwidth]{./bbcp_vs_file_80_1MB.pdf}\\ \small (c) Small loads 60\% fault time & \small (d) Small loads 80\% fault time \\ \end{tabular} % \vspace{0.1in} \caption{\small Recovery time analysis of FileLogger at varying fault timing for small workloads. } \label{fig:bbcpvsFL_small} % \vspace{0.1in} \end{figure*} %In this section, Recovery times with all object based fault tolerance mechanism and methods are compared with that of bbcp data transfer tool. Where we set LADS recovery time as the baseline for our experiments. As resume operation is not supported in LADS, LADS has to transfer all the objects of the dataset upon resuming from faults. From the experimental results shown in Figure~\ref{fig:bbcpvsFL_big},~\ref{fig:bbcpvsFL_small} and ~\ref{fig:ft_rec}, the later the fault occurs, the higher the recovery time is. %\cred{Why do you only show recovery times of the file logger, not other two loggers? } %{\color{blue} Already mentioned in next paragraph} Our aim is to minimize the impact of recovery time on the fault point. As per our logging mechanism, we delete the log file entries of the logical files, which are successfully transferred to the sink end. Due to this at any point of time, we are left with only those files which are currently being progressed. %\cred{This statement is only valid for file logger, whereas the log files last to exist during execution of data transfer for other logger implementations. } %{\color{blue} Here i am describing about the logical file entries and hence this is valid for all mechanisms. In case of universal and transaction logger, file entries are deleted from the list} %\cred{I understand now. Ignore my comment. } The amount of logs to be parsed to retrieve the objects which are successfully synchronized at the sink end PFS will not depend on the fault point. Recovery time of File logger mechanism at varying fault points for both big and small workloads is as shown in Figure~\ref{fig:bbcpvsFL_big} and~\ref{fig:bbcpvsFL_small}. For other Transactional and Universal FT mechanisms, similar results were observed. We only show the results for File logger in this paper. %\cred{I think the above paragraphs could be better organized and written. Can you read writing till here and improve the organization of your writing?} %{\color{blue} I had organized as, Simulation environment, baseline LADS and our recovery approach. Please suggest how this can be reorganized} %\cred{} \subsubsection{Big Workloads} \label{sec:bigwork} In case of file logger mechanism, the recovery times for all fault tolerance methods exhibit similar recovery times irrespective of the fault points (Refer to Figure~\ref{fig:bbcpvsFL_big}). Though the recovery time is much lower than LADS, all the methods of file logger mechanism consume higher recovery times than bbcp. %\cred{Do you know why bbcp recovery time is very small?} %{\color{blue} As bbcp FT is based on file offset, its recovery time is much less than that of file logger. %} %{\color{green} This is expected as in file logger mechanism, each logical file to be transferred is associated with one log file and while writing the logs to file, we just append the completed object index at the end of logger file. Due to this, while retrieving the completed object information, an additional search overhead is involved. %} %\cred{You explain here the recovery time of file logger is not as small as bbcp but, you never discussed why the recovery time of the file logger is higher than that of bbcp.} %{\color{blue} explanation is in green please check.} For transaction and universal loggers, the recovery time overhead of big workloads is negligible. This is also expected, as the completed objects information is sorted as per object index before writing to the logger file. As mentioned in Section~\ref{subsec:perf_comp}, we are using intermediate lists, which maintain the completed objects information of all files being transferred, by sorting based on object index. \subsubsection{Small Workloads} In contrast to the big loads, bbcp tool consumes much higher transfer time for smaller workloads than LADS. %{\color{green} Due to this, the recovery time overhead of FT-LADS is not directly comparable with bbcp. For quantitative comparison, percentage of recovery time relative to each method is calculated. At all given fault points, bbcp exhibit 5\% to 7\% recovery time overhead. Whereas, all the proposed FT methods experience around 12\%-14\% overhead. %} %But from the experimental results, bbcp's recovery time for small workloads is much higher than big workloads. %\cred{Why does bbcp's recovery time show higher for big workloads than for small workloads?} %{\color{blue} bbcp show higher recovery time for small workloads. This might be due to the number of files in the workload. } %\cred{Kasu, referring to the results of Fig7, in particular, (a), if I only compare recover time, I can see clearly, bbcp recovery time is smaller than any methods of FT-LADS, but bigger than LADS's recover time. LADS's recovery time is highest. These results make sense to me. On the other hand, for small workloads, LADS, FT-LADS's recovery times are very invisible. It's because compared to runtime of bbcp, their runtimes are quite small, so their recovery times are very invisible. File. But, why is bbcp's recovery time visibly big? You need to explain why bbcp's recovery time shows higher than others. } %{\color{blue} Professor. I will check this and update you. I was thinking on fault, all the files need to be parsed again. As in small workloads scenario the number of files are high, bbcp is taking long time. I cross checked the results, but there was no mistake. I will check again for the cause.} %With this we can assume that bbcp's fault tolerance method best suits for big work loads. %\cred{Don't you have better evidence than this assumption?} %{\color{blue} This was conclusion from the experimental results.} %\cred{I don't agree with your conclusion. Better to delete this conclusion. } Our small workload consists of files whose size is of 1MB and this is matching our transfer unit size. Due to this, a file transfer state can be either completed or transferred upon recovery from fault. So, there won't be any log files which need to be parsed upon fault, and hence, proposed object based logger mechanisms just determine which files are already completed and start transferring the remaining files. As a result, we can conclude that with the proposed object based fault tolerance mechanisms, the recovery time overhead will not come into the picture. %\cred{Why does bbcp's recovery time show visible portion whereas file logger methods do not?} %{\color{blue} As bbcp overall transfer time is much higher, recovery time is significant. whereas LADS case transfer time is less than 10sec. So recovery time is not siginificant. } %\cred{What if you calculate percentage of recovery time? Out of 10 sec, how does does it take for recovery? 10percent or 20 or so? Is this portion larger or smaller than that of bbcp's recovery? } %{\color{blue} I had calculated the recovery time \% and bbcp is taking around 6\% whereas FTLADS is taking around 10\%-12\%} %\cred{Great. bbcp's recovery time portion is smaller than FTLADS. It makes sense to me. Why don't you show these percentage numbers to compare the results? } \begin{figure}[!t] \centering \begin{tabular}{@{}cccc@{}c@{}} \includegraphics[width=0.4\textwidth]{./ftlads_recovery_80_1GB.pdf} & \hspace{0.3in} \includegraphics[width=0.4\textwidth]{./ftlads_recovery_80_1MB.pdf} \\ \small (a) Big loads 80\% fault time & \small (b) Small loads 80\% fault time \\ \end{tabular} \caption{\small Recovery time analysis of FT Loggers at 80\% fault timing. The 99\% confidence intervals are shown in error bar.} \label{fig:ft_rec} %\vspace{0.1in} \end{figure} In Figure~\ref{fig:ft_rec}, we have shown the recovery time comparison among the proposed fault tolerance mechanisms, considering 80\% fault point as a reference, for both big and small workloads. As shown in Figure~\ref{fig:ft_rec}, we can observe that for big workloads, file logger mechanism exhibits higher recovery time than other proposed FT mechanisms. Whereas, for small workloads, the recovery overheads for all mechanisms and methods are similar as shown in Figure ~\ref{fig:ft_rec} (b). %\cred{In Fig8, do they only show ** recover times **? I don't understand why file logger show higher recovery time than other loggers? For small loads, why char, bits, and bit64 show higher recovery time whereas they do not for big loads? Analysis for the results of Fig8 is incomplete.} %{\color{blue} For big workloads, file logger will have higher recovery time and the same is explained in green font in the above section.} %\cred{Justify your observation here again} %{\color{blue} For small workloads, i missed to update the graph after my last week's modifications. I had updated the graph now} %\cred{adjust y-scale in Fig8(b) so that bars are visible} %\cred{Ok.} From Figure ~\ref{fig:ft_rec} (a) and ~\ref{fig:ft_rec} (b), we can observe that Universal logger mechanism exhibits lower recovery times upon fault. Also, among all the FT methods, bitbinary methods (Bit8 and Bit64) have minimal recovery overhead compared with the other FT methods. %\cred{You should explain why file logger design shows higher recover time!} %{\color{blue} I had given the probable reason in section, ~\ref{sec:bigwork}.} Based on our evaluation results, file logger mechanism shows minimal impact on the performance while logging the completed objects information. Whereas, universal logger is superior to other mechanisms with respect to recovery times upon fault. Also, among the proposed FT methods, bitbinary methods (Bit8 and Bit64) have minimal space overhead and prove to have comparably lower recovery times among all the proposed FT mechanisms. Hence, conjugating LADS with universal object based FT mechanism and bitbinary FT methods will improve data transfer performance in faulty environments. \begin{comment} %\cred{Kim: wrong title! What about recovery time analysis?} %For evaluating recovery time, bbcp tool has been used as a benchmark. %\cred{Kim: bbcp is not a benchmark. It's another data transfer tool.} We have analyzed our recovery time for WL$_{S}$ and WL$_{B}$ and compared the same against bbcp data transfer tool. For effective evaluation of recovery time of proposed fault tolerance methods, we %had %\cred{have} have created a simulation environment in which we %had have generated hardware faults after transferring 20\% , 40\%, 60\%, 80\% of total data size. %\cred{Kim: if you consider software fault, why do you have to clear out page cache? In case of hardware failure (reboot), page cache will be cleared. Your experiment assumes hardware failure, I think.} %\cred{Kim: specify if you make source side failure or sink side. And why you pick one side failure. } As faults can be occurred at any end of the transfer, we can simulate the faults at either source or sink. However, for the purpose of our experiments, we have executed this simulation in the source end. We %had have %\cred{Kim: I suggest using `` present perfect'' or ``past'' form in writing.} used the environment described in Section~\ref{sec:expr} %\cred{Kim: what section are you referring to???} for measuring recovery time for both bbcp and FT-LADS. On these hosts, LADS uses CCI’s Verbs transport, which natively uses the underlying InfiniBand interconnect. Whereas, bbcp uses the IPoIB interface which supports traditional sockets. \cred{Kim: This has been already mentioned.} \textcolor{blue}{i had removed the above and included here} In LADS, varying the number of I/O threads, maximizes CPU utilization on the data transfer node. However, bbcp uses configurable window size and number of streams to improve the performance. In our test, we %had %\cred{ %Kim: you need to mention why you pick four I/O threads in LADS and 2 tcp streams and such in bbcp. You need to justify this setting does not change your following observation from Fig4. %} Based on the experimental results presented in LADS paper, \cite{kim:fast15}, LADS data transfer performance increase linearly with the number of i/o threads. Whereas, bbcp has less impact while increasing the number of tcp streams. To have optimal performance comparison between two, we have fixed FT-LADS to use $4$ I/O threads and bbcp to use $2$ tcp streams and window size of 8MB. Our experiments were designed to calculate the transfer time before and after fault. Based on these times and using \ref{equ1} we have estimated the recovery time. %\cred{Kim: I suggest moving this equation and illustration to Sec III. C (Resuming Failed Transfers)? Then, here, you may just refer equation parameters in explaining and analyzing results.} \begin{figure}[!t] \centering \begin{tabular}{@{}c@{}c@{}} \includegraphics[width=0.24\textwidth]{./images/20_1GB}& \includegraphics[width=0.24\textwidth]{./images/20_1MB}\\ \small (a) Big loads 20\% fault time & \small (b) Small loads 20\% fault time\\ \end{tabular} % \vspace{0.1in} \begin{tabular}{@{}c@{}c@{}} \includegraphics[width=0.24\textwidth]{./images/40_1GB}& \includegraphics[width=0.24\textwidth]{./images/40_1MB}\\ \small (c) Big loads 40\% fault time & \small (d) Small loads 40\% fault time\\ \end{tabular} % \vspace{0.1in} \begin{tabular}{@{}c@{}c@{}} \includegraphics[width=0.24\textwidth]{./images/60_1GB}& \includegraphics[width=0.24\textwidth]{./images/60_1MB}\\ \small (e) Big loads 60\% fault time & \small (f) Small loads 60\% fault time\\ \end{tabular} % \vspace{0.1in} \begin{tabular}{@{}c@{}c@{}} \includegraphics[width=0.24\textwidth]{./images/80_1GB}& \includegraphics[width=0.24\textwidth]{./images/80_1MB}\\ \small (g) Big loads 80\% fault time & \small (h) Small loads 80\% fault time\\ \end{tabular} % \vspace{0.1in} \caption{\small Recovery time analysis at varying fault timing.} \label{fig:bbcpvsft} \vspace{0.1in} \end{figure} Figure \ref{fig:bbcpvsft}, represents the analysis of FT-LADS recovery time with bbcp tool for a simulated fault at 20\%, 40\%, 60\% and 80\% data transfer points. For easy understanding of the recovery time, the graph was represented using stacked histogram, and the recovery time has shown between the consumed time before fault and consumed time after fault. Figure \ref{fig:bbcpvsft}(a) depicts the recovery time analysis for WL$_{B}$ at 20\% fault time. All of the proposed fault tolerance methods are experiencing a constant recovery time, where as we have observed that bbcp net recovery time in this case is negative and hence was not shown in the graph. \cred{Kim: what can the recovery time negative? Maybe evaluation error???} Figure \ref{fig:bbcpvsft}(b) show the recovery time analysis for WL$_{S}$ at 20\% fault time. WL$_{S}$, we have observed that recovery time of FT-LADS is very much comparable with bbcp. All other graphs in Figure\ref{fig:bbcpvsft} represent the recovery time comparison between bbcp and FT-LADS at 40\%, 60\% and 80\% fault points for both WL$_{S}$ and WL$_{B}$ Even though there is slight performance degrade with WL${S}$, overall performance and recovery time is very much comparable with LADS and bccp tools respectively. With these experiments, we have successfully achieved the our experimental goals. \cred{Kim: Kasu, mention why recovery time in Fig4(a-b) is decreasing as fault time happens later. Did you see the same observation from small workloads? If not why? If so why again? Analyzing why recover time is high in portion to other times in small workloads, which is not true in big workload? I'm saying, why recovery time in small workload result is dominant (around 20-30\%)? You should compare with results when no failure in LADS. Basically that's the overhead to be counted in failure in LADS by fault tolerance mechanism. } \end{comment} %%%%%%%%%%%%%%%%%%%%%%%% \begin{comment} content... To measure the performance overhead of each logging mechanism, we implement each method using a single logging thread and evaluate space overhead and file system logging time of each mechanism/method.} {The space overhead of each method is also computed based on space requirement for each log entry in log files, and its associated data structure overhead.} All files can be segmented to 1MB blocks. Based on the file size distribution mentioned in LADS~\cite{kim:fast15_short}, we consider two representative file sizes: small (1MB) and big (1GB). \begin{figure}[!t] \centering \begin{tabular}{@{}c@{}c@{}} \includegraphics[width=0.24\textwidth]{./Source_GnuPlot/expr_space_avg}& \includegraphics[width=0.24\textwidth]{./Source_GnuPlot/expr_time_avg}\\ \small (a) Log File Size & \small (b) Transfer Time for One File\\ \end{tabular} \vspace{-0.1in} \caption{\small Comparison of %proposed all fault tolerance methods. The 99\% confidence intervals are shown in error bar. } \label{fig:FTCompare} \vspace{-0.2in} \end{figure} Figure~\ref{fig:FLSmall} shows space and performance comparison for both small (1MB) and large size files (1GB). %As described in section \ref{sec:FL}, file logger fault tolerance has been compared for various methods of logging. We have analyzed the file logging fault tolerance method for files of size as described above. \ref{fig:FLSmall}. %From the above Figure \ref{fig:FLSmall}, %Figure~\ref{fig:FLSmall} shows that For small size files, space and performance optimization can be achieved using encode method of logging. On the other hand, for large files, space optimization can be achieved using %Bitbinary Bit binary method of logging, however, the performance overhead is significant in the Big binary method. %Choosing one of the method depends on the needs of the application. %Similar experiments %has %\cb{have} been conducted on other type of proposed fault tolerance methods -- transaction logger and universal Logger. Figure~\ref{fig:FTCompare} shows the comparison of the space and execution time for all the three proposed fault tolerance mechanisms. %For simulation, For workloads, we used 1GB big files. %we used $2,600$ 1~GB files.} %we have considered $2600$ files all are of equal size, $1GB$. %The graph depicted %below %shows the average space and time consumed for one file. From the figure, it is evident that file logger mechanism of fault tolerance consumes lowest space as compared to the other fault tolerance mechanisms. Whereas, the universal logger mechanism is efficient with respect to logging time. % the amount of time consumed. %\cred{Kim: need more explanations about Figure~\ref{fig:FTCompare}.} \end{comment} \section{Related Work} \label{sec:related} %\cred{This section needs improvement.} To meet the needs of big data transfers, prior studies have performed on the design and implementation of bulk data movement frameworks ~\cite{Allcock:2005:GSG:1105760.1105819, bbcp, xdd:brad, sc12:rftp, Ren:sc13, Subramoni:2010:HPD:1844765.1845179, 13:vallee:ndm, DBLP:journals/tpds/LiRYJ17}. GridFTP~\cite{Allcock:2005:GSG:1105760.1105819}, which is an extended version of the standard File Transfer Protocol (FTP), provides high speed, reliable, and secure data transfer. The striping feature in GridFTP enables the support for multi-host to multi-host transfers. But this tool does not try to schedule the data transfer based on the underlying object locations. \textit{bbcp}~\cite{bbcp} is another data transfer tool which uses multiple streams for transferring large datasets. It uses a file based approach, which transfers the whole file data sequentially. XDD~\cite{xdd:brad} optimizes the disk I/O performance by enabling file access with direct I/Os and using multiple threads for parallelism, and varying file offset ordering to improve I/O access times. RAMSYS~\cite{DBLP:journals/tpds/LiRYJ17}, a resource-aware high-speed data transfer software, utilizes a multi-stage end-to-end data transfer pipeline, where each stage is fully resource-driven and implements a flexible number of components using predefined functions, such as storage I/O, network communication, and request handling. RAMSYS relies on the asynchronous paradigm to maximize the concurrency of components and thereby offers improved scalability and resource utilization in modern multi-core systems. All these tools are useful for moving large data faster and secure from source host to remote host over the network, but none of them tries to schedule based on the underlying object locations because they do not consider storage contention. %{\color{blue} I tried to find some information on FT support in XDD and RAMSYS but i couldnt find any information on this. So, here i just described about GridFTP and bbcp} %\cred{Then write that XDD and RAMSYS did implement fault tolerance in next paragraph.} %{\color{blue} This i am not sure professor. As i could not find any reference for FT in XDD and RAMSYS. If you want me to specify like below, i can write. But there is no reference.Like, bbcp and GridFTP, XDD and RAMSYS also implement file based logging mechanisms for FT.} %Checkpoint based file offset logging is common method of fault tolerance in most of the existing data movement tools. Here we are considering, popular data trasfer tools, like \textit{bbcp}~\cite{bbcp} and \textit{GridFTP}~\cite{Allcock:2005:GSG:1105760.1105819} as a reference. Both these tools have the support of handling the faults occurred during transfer and resuming from the failed transfers upon recovery. Another important aspect of these data movement frameworks is to resume the data transfer upon faults during data transfer. GridFTP tool supports fault tolerance using restart markers (checkpoints). While transferring data, GridFTP server automatically sends restart markers to the client. If the transfer has a fault, the client may restart the transfer by providing the markers received. The server will restart the transfer from the point where it left off based on the markers. GridFTP's Reliable File Transfer (RFT) service provides an interface to write the restart markers to a database so that it can survive a local fault. \textit{bbcp} tool employs fault tolerance mechanism based on checkpoint record. Upon initiating a new transfer, \textit{bbcp} tool checks if checkpoint record of file being transferred exists or not. If record does not exist, it checks the target file attributes like name, size, etc. If they are identical with the source file attributes, then \textit{bbcp} assumes that the file transfer completed successfully and skips the transfer. If file attributes are different, then it initiates a new transfer by creating a checkpoint record and transmit all the source bytes to the target. Upon successful completion, it erases the checkpoint record. If checkpoint record exists, then it resumes the transfer by appending all untransmitted bytes to the target. XDD and RAMSYS tools did not implement fault tolerance, because they did not know which data needs to be transmitted upon fault. As all the aforementioned bulk data movement frameworks transfer the logical file data sequentially, it is possible to resume transfers using checkpoint based restart marker or offset record. Checkpoint based fault tolerance methods are light-weight and also possible to resume transfer from restart marker or offset record without delay. Our work focuses on entirely different scenario from the prior fault tolerance studies. Our work focuses on supporting resume functionality upon fault when the workload is transferred as objects rather than files, by exploiting the underlying storage architecture. Since a logical file is striped over multiple OSTs, it is possible to transfer one logical file's objects in random order. While, above mentioned checkpoint based restart marker or offset record is not sufficient to resume the transfer upon fault, our work proposes novel methods to handle fault tolerance in object based big data transfers. In our proposed object based fault tolerance mechanisms, objects which are successfully written to sink PFS are marked as successful and we update the information of the object in the logger file. Upon successful completion of all the objects of one logical file, the log information corresponding to the file will be erased. If there is any fault during the transfer, the proposed mechanisms search for the completed objects and schedule only those objects which are not transferred previously. As in object based fault tolerance mechanism, it needs to log all the objects of a file. This involves access to asynchronous filesystem API which causes processing overhead. It also results in space overhead as all the objects information is logged to the logger file. It involves an additional overhead to retrieve the completed objects information from the logger file for resuming the transfer upon recovery from fault. Our solution proposes methods to overcome processing, space and recovery time overheads. %\cred{Kim: Kasu, can you write briefly how your logging idea is different from the approaches employed in the existing tools such as bbcp? Few sentences will surffice.} \section{Conclusion} \label{sec:conc} \begin{comment}{To support fault tolerance in LADS, we have implemented six different fault tolerance methods: Char, Int, Enc, Binary, Bit8 and Bit64. Also, we had evaluated the performance impact of FT-LADS on LADS and from the experimental results, we had observed that FT-LADS doesn't have impact on the total data transfer time but has a CPU overhead of around 10\%. We had also compared the fault recovery of FT-LADS with bbcp and found that FT-LADS recovery time is very much comparable with bbcp recovery time. Considering the file logger, log file meta data overhead, in future we are going to extend our work with transaction and universal logger. In future we can extend our work to support data integrity check at the sink end. Also, based on the environmental conditions, dynamically choose the fault tolerance mechanism such that it will have minimal recovery time, incase of errors. \end{comment} LADS data transfer tool with its layout-aware and OST congestion-aware algorithms, outperforms existing data transfer tools with respect to the data transfer rate. However, absence of fault tolerance support results in the data retransmission upon fault. %and hence congestion along the data transfer path. %\cred{why does the congestion happen?} %{\color{blue} Professor, data retransmissions to compensate failure cause congestion} %\cred{Although unnecessary traffic can be generated, congestion does not necessarily occur. I would suggest to not mention congestion.} As LADS employs object level scheduling algorithms, objects of one logical file may be transferred out of order. Due to this, the fault tolerance mechanisms based on logging file offset, are not suitable for LADS. In this work, we have implemented object based fault tolerance mechanisms which can handle the out-of-order nature of object transmission. Depending on the number of logger files generated per dataset, we have proposed three different object logger mechanisms, \textit{File logger, Transaction logger and Universal logger}. Also in order to reduce the space overhead of logging, we have proposed six different fault tolerance methods: Char, Int, Enc, Binary, Bit8 and Bit64. We have evaluated the performance overhead of fault tolerance on LADS and concluded that proposed object based logging mechanisms do not impact the LADS data transfer performance. In order to evaluate the recovery time overhead of proposed object based fault tolerance mechanisms, we have created simulation environment to generate faults at 20\%, 40\%, 60\% and 80\% points of data transfer. %\cred{\sout{With this, we have simulated faults and compared the recovery time overhead with bbcp data transfer tool. } %\sout{From the evaluation results, we have concluded that the recovery time is independent of the fault point. }} %Also, From our evaluation results, we have observed that the recovery time in file logger mechanism exhibits $2$ times higher than bbcp, whereas the recovery times in transaction and universal logger mechanisms were considerably smaller than bbcp. %\cred{When I finish reading conclusion till this, it seems file logger is the worst option, because while performance overhead of all three loggers is negligible, the recovery time of the file logger is said the worst. Is this what you want to claim? I don't think so. You wanted to say that out of three loggers, normal operation performance overhead of the file logger is the most minimal however, the recovery time is worse than other transactional or universal loggers. But as I said, we could claim fault doesn't often happen, so performance overhead minimization during normal operation is more important than recovery time and even if the file logger is the worse out of three, but still the recovery time of the file logger is tolerable.} %\cred{ Can you write that even though FT-LADS's recovery time is x times higher than bbcp but, it's not significant. } %With this, we can conclude that LADS in conjunction with proposed object based fault tolerance mechanisms will improve the data transfer performance in faulty environments by avoiding already transferred objects, upon recovery from fault. %\cred{This sentence needs to be modified. } To conclude, \textit{File logger} mechanism has minimal impact on logging the completed objects. %\cred{Can you read this sentence and beginning parts in conclusion? This sentence is not well connected with prior story.} Whereas, universal logger mechanism combined with bitbinary methods (Bit8 and Bit64) has a minimum overhead with respect to space and recovery times. Hence, with the proposed fault tolerance mechanisms, LADS can experience improved recovery performance upon fault during transfer. %In this study, we focused only on the errors occured while transfering the data from source to sink. %In future, we will extend our work to support, data integrity check to avoid orruption of data while transfering from one node to another or while storing the data to PFS. Also, currently LADS consider the data transfer between two nodes. In future we will extend this work to support multi-node parallelization. %Several future direction has been identified: (i) integrating the proposed mechanisms with LADS data transfer framework, (ii) adding an write confirmation message at sink host to source host, and (iii) developing asynchronous loggers at source host.} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{comment} \section{Conclusion} \label{sec:Conc} To minimize the data integrity and performance issues due to errors occurred in end-to-end path of data transfer, three methods of fault tolerance methods has been proposed. %File Logger, Transaction logger and universal logger. Also, comprehensively evaluated the space and write performance overhead of proposed fault tolerance methods. Our evaluation results showed that proposed fault tolerance methods will have very minimum impact on LADS with respect to space and time overhead. \end{comment} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% Kim: we don't need future in this submission %%% %\section{Future Work} %\label{sec:Future} %Analysis of search complexity to retrieve the completed block information in all the logging methods specified in this paper. Adaptive implementation of Universal, Transaction and File logger methods to reduce the search overhead and also to improve the overall performance of the data transfer. Integrating proposed fault tolerance method with LADS data transfer frame work. \section*{References} \bibliography{zs} \begin{comment} \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/Kasu.jpg}} \noindent{\bf Preethika Kasu} is currently pursuing MS degree in the department of software and computer engineering at Ajou University, South Korea. Her research interests include fault tolerance, distributed file and storage, parallel I/O and high-performance computing.\\ \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/Taeuk.jpg}} \noindent{\bf Taeuk Kim} is currently pursuing MS degree in computer science and engineering department at Sogang University, South Korea. He received BS degree in the department of computer science and engineering from Sogang University in 2017. His research interests include parallel and distributed file system, and memory power consumption.\\ \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/Prof_Kim.jpg}} \noindent{\bf Youngjae Kim} received the BS degree in computer science from the Sogang University, Korea in 2001, MS degree from KAIST, South Korea in 2003, and PhD degree in computer science and engineering from the Pennsylvania State University in 2009. He joined Sogang University as a faculty member in the department of Computer science and engineering in 2016. Before joining Sogang, he was a faculty member in the department of software and computer engineering at Ajou University, South Korea in 2015-2016. Prior to joining Ajou University, he was a research staff member in the Oak Ridge National Laboratory in 2009-2015. His research interests include distributed file and storage, parallel I/O, operating systems, emerging storage technologies, and performance evaluation.\\ \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/jhum.jpg}} \noindent{\bf Jung-Ho Um} is a senior researcher at Korea Institute of Science and Technology Information (KISTI). He received his PhD in Computer Engineering from Chonbuk National University, Korea. His current research interest includes database, information retrieval, distributed computing, and multi-dimensional data analysis.\\ \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/gspark.jpg}} \noindent{\bf Kyongseok Park} is a senior research scientist and R\&D leader in scientific data technology laboratory at Korea Institute of Science and Technology Information (KISTI). His current research interests are distributed computing, high-performance computing, numerical analysis and machine learning.\\ \parpic{\includegraphics[width=1in,clip,keepaspectratio]{photos/scott.jpg}} \noindent{\bf Scott Atchley} received BS degree in business administration and MS degree in computer science from the University of Tennessee in 1987 and 2002, respectively. He joined Oak Ridge National Laboratory as a HPC systems engineer in 2011. He is the team lead for System Architecture, Resilience, and Networking in the Technology Integration Group within ORNLs National Center for Computational Science. His research interests include high-performance interconnects and their interfaces, system architectures, and multi-level memories. Prior to joining ORNL, he was a member of the technical staff at Myricom and a research leader at the University of Tennessee.\\ \end{comment} \begin{comment} \newpage \section*{Highlights} \begin{itemize} \item{Objects of one logical file are transferred out of order due to the object level scheduling algorithms of LADS data transfer tool} \item{Logging file offset based fault tolerance methods are not suitable with LADS data transfer tool} \item{Proposed object logging based fault tolerance methods efficiently handle the log file space overhead} \item{Proposed fault tolerance mechanisms to handle the number of logger files per dataset} \item{Recovery time from fault is comparable with sequential data trasnfer tools, like bbcp} \item{Negligible impact on LADS data transfer rate upon fault during data transfer} \end{itemize} \end{comment} \end{document}}\end{comment}}
\caption{The architecture of NMT with adversarial stability training. The dark solid arrow lines represent the forward-pass information flow for the input sentence $\mathbf{x}$, while the {\color{red}red dashed arrow lines} for the noisy input sentence $\mathbf{x}^{\prime}$, which is transformed from $\mathbf{x}$ by adding small perturbations. %represents the information flow of the forward pass for $\mathbf{x}$, while the red arrow line displays the information flow for $\mathbf{x}^{\prime}$. The sentence $\mathbf{x}$ is first transformed to $\mathbf{x}^{\prime}$ by adding small perturbations. $\mathbf{h}_{\mathbf{x}}$ and $\mathbf{h}_{\mathbf{x}^{\prime}}$ are representations encoded by the encoder respectively for $\mathbf{x}$ and $\mathbf{x}^{\prime}$ }
\caption{Comparison of extractive, abstractive, and our unified summaries on a news article. The extractive model picks most important but \textcolor{myblue}{\textbf{incoherent or not concise}} (see blue bold font) sentences. The abstractive summary is readable, concise but still \textcolor{red}{\textit{loses or mistakes some facts}} (see red italics font). The final summary rewritten from \underline{fragments} (see underline font) has the advantages from both extractive (importance) and abstractive advantage (\textcolor{mygreen}{\textbf{coherence}} (see green bold font)).}
\caption{\emph{Example} for \Cref{alg:main}. The $2$ update flow pairs are \textcolor{red}{red} and \textcolor{blue}{blue}, each of demand $1$. The active edges of the respective colors are indicated as \emph{solid lines} and the inactive edges are \emph{dashed}. Each edge in the flow graph is annotated with its current load (\emph{top}) and its capacity (\emph{bottom}). We start by identifying the \textcolor{blue}{blue} and \textcolor{red}{red} blocks. For \textcolor{red}{red} there is exactly one such block \textcolor{red}{$r_1$}, since \textcolor{red}{$R^o$} and \textcolor{red}{$R^u$} only coincide in $s$ and $t$. The \textcolor{blue}{blue} flow pair on the other hand omits two blocks \textcolor{blue}{$b_1$} and \textcolor{blue}{$b_2$}: \textcolor{blue}{$B^o$} and \textcolor{blue}{$B^u$} meet again at $w$ and at $t$. We observe that $\textcolor{blue}{b_2}$ can only be updated after $\textcolor{red}{r_1}$ has been updated; similarly, $\textcolor{red}{r_1}$ can only be updated after $\textcolor{blue}{b_1}$ has been updated. An update sequence respecting these dependencies can be constructed as follows. We can first prepare the blocks by updating the following two out-edges which currently do not carry any flow: $(w,\textcolor{red}{red})$, $(u,\textcolor{blue}{blue})$, and $(v,\textcolor{blue}{blue})$. Subsequently, the three blocks can be updated in a congestion-free manner in the following order: Prepare the update for all blocks in the first round. Then, update $\textcolor{blue}{b_1}$ in the second round, $\textcolor{red}{r_1}$ in the third round, $\textcolor{blue}{b_2}$ in the fourth round. }
\caption{(Watts vs. Seconds) Canonical shapes found by clustering three-second intervals of training data. Note that the four shapes characterize {\color{red}high power}, {\color{blue} low power}, {\color{darkolivegreen} falling power}, and {\color{teal} rising power}. }
\caption{(Watts vs. Seconds) Canonical shapes found by clustering three-second intervals of training data. Note that the four shapes characterize {\color{red}high power}, {\color{blue} low power}, {\color{darkolivegreen} falling power}, and {\color{teal} rising power}. }
\caption{An overview of the {\it JointEmbedding} model. The two embedding components \imemb (in {\color{myyellow} yellow}) and \sentemb (in {\color{red} red}) are shown on the left while the \sentgen (in {\color{gray} grey}) is on the right. }
\caption{An overview of the {\it\mymodel} model. The \termgen network (in {\color{green} green}) is shown in the lower left. The \langgen network is in the upper right (in {\color{blue} blue}) . }
\caption{Molecular Dynamics Simulations (adapted from Ref.~\cite{Schachoff-etal:2015}). (a) Snapshot of a heated Janus particle with wetting parameters $c_\mathrm{gs}=2,~c_\mathrm{ps}=1$ on the two hemispheres; coloring indicates the measured kinetic energy from which the continuum temperature field $T(r,\theta)$ depicted in (b) is deduced. (c) Lab frame $(x,y,z)$ and (co-rotating) body frame $(\bar{x},\bar{y},\bar{z})$. (d) Solvent temperature $T(a,\theta)$ in a thin shell of thickness $0.5\sigma$ around the heated Janus particle, with its cap maintained at temperatures $T_\mathrm{p}=1.20 \epsilon/\kb $ ({\Large \color{myblue} $\bullet$},{\Large \color{myblue} $\circ$}); $1.50 \epsilon/\kb$ ({\small \color{mypurple} $\blacksquare$},{\small \color{mypurple} $\square$}) and $2.00 \epsilon/\kb$ ({\color{myokker} $\blacktriangle$},{\scriptsize \color{myokker} $\triangle$}), for wetting parameters $c_\mathrm{gs}=2$, $c_\mathrm{ps}=1$ and $c_\mathrm{gs}=1$, $c_\mathrm{ps}=2$ (inset). Solid lines represent fits by the series expansion from Eq.~(\ref{eq:heat_eq_soln}) and dashed lines solutions of the heat equation with the appropriate temperature-dependent thermal conductivity \cite{Chakraborty:2011}, both truncated after $n=3$ (which causes spurious oscillations). }
\caption{Mean-square displacements (a) and time-dependent diffusivities (b) along (filled symbols) and perpendicular to (empty symbols) the propulsion direction of a heated Janus particle ($c_\mathrm{gs}=2$ and $c_\mathrm{ps}=1$) in the particle frame. The temperatures of the hot cap are $T_p=1.10~\epsilon/\kb$~({\color{myblue}{\Large $\circ$, $\bullet$}}), $1.50~\epsilon/\kb$~({\color{mypurple}{\small $\square$,$\blacksquare$}}) and $2.00~\epsilon/\kb$~({\color{myokker}{$\vartriangle$,$\blacktriangle$}}). Adapted from Ref.~\cite{Schachoff-etal:2015}.}
\caption{Passive and active propulsion velocities for homogeneous beads and Janus beads. (a) Passive phoretic velocities for a homogeneous particle with $c_\mathrm{ps}=c_\mathrm{gs}=1$ ({\Large \color{myblue} $\bullet$}) and $c_\mathrm{gs}=c_\mathrm{ps}=2$ ({\small \color{mypurple} $\blacksquare$}) and a Janus particle with $c_\mathrm{ps}=2, c_\mathrm{gs}=1$ ({\color{mygreen} $\triangledown$}) and $c_\mathrm{ps}=1, c_\mathrm{gs}=2$ ({\large \color{myokker} $\diamond$}) (inset). (b) The component $\bar T B_1/(a+s)$ of the surface temperature gradient that causes the propulsion, as measured in fluid shells of various thicknesses $s$ around the Janus bead. The active and passive propulsion velocities are the same for each heating power/temperature gradient, therefore the measured $\bar T B_1/(a+s)$ should also coincide for Eq.~(\ref{eq:mueff}) to hold. Lines guide the eye in inferring the corresponding effective boundary layer thickness $s$ (gray bar). }
\caption[Evolution of a passing wave]{Evolution of a passing wave ($\mu=0.1$, $\varepsilon=0.2$) over a small obstacle ($\beta=0.3$, $c_{fric}=0.5$); {\color{red}red} curve is the passing wave, {\color{green}green} curve is the reference soliton for flat bottom}
\caption[Evolution of a breaking wave]{Evolution of an approaching large wave ($\mu=0.25$, $\varepsilon=0.35$ and breaking when reaching the obstacle ($\beta = 0.5$, $c_{fric}=0.5$); {\color{red}red} curve is the approaching wave, {\color{green}green} curve is the reference soliton for flat bottom}
\caption[Evolution of a passing wave over a sliding object]{Evolution of a passing wave ($\mu=0.1$, $\varepsilon=0.2$) over a larger sliding obstacle ($\beta=0.4$, $c_{fric}=0.001$); {\color{red}red} curve is the passing wave, {\color{green}green} curve is the reference soliton for flat bottom}
\caption[Evolution of a surging wave]{Evolution of an approaching wave ($\mu=0.25$, $\varepsilon=0.25$) over an almost perfectly sliding large solid ($\beta=0.5$, $c_{fric}=0.001$); {\color{red}red} curve is the approaching wave, {\color{green}green} curve is the reference soliton for flat bottom}
\caption{\textbf{Results of our multi-view model with $f$ and $g$ and other variants}. In the table, ``Avg of STS tasks'' refers to the mean Pearson's score on five STS tasks; ``Avg of SICK-R, STS-B'' refers to the mean Pearson's score on Sick-Entailment and STS-Benchmark as they both require the same feature engineering methods proposed in \cite{Tai2015ImprovedSR}; ``Avg of Binary-CLS tasks'' refers to the mean accuracy on five sentiment analysis tasks; $en(\cdot,\cdot)$ stands for an ensemble of two representations. The arrow indicates the performance boost (\textcolor{ForestGreen}{$\uparrow$}) or drop (\textcolor{red}{$\downarrow$}) relative to the same part in our model, e.g., \textcolor{red}{$\downarrow$17.7} indicates the performance of $\mathbf{z}^f$ in multi-view with $f_1$ and $f_2$ is 17.7 point lower than that of the $\mathbf{z}^f$ in our multi-view model with $f$ and $g$. Better view in colour. }
\caption{Visualization of interpretable filters in the explainer and ordinary filters in the performer. As discussed in \cite{Interpretability}, the top conv-layer of a CNN is the more likely to represent object parts than low conv-layers. We visualized and compared filters in the top conv-layer of the performer and interpretable filters in the \textit{conv-interp-2} layer of the explainer. We used \cite{CNNSemanticDeep} to estimate the RF\textcolor{red}{{\protect\footnotemark[3]}} of neural activations to illustrate a filter's semantics. Interpretable filters are much more semantically meaningful than ordinary filters. Please see the appendix for more results.}
\caption{Visualization of clusters on the Amazon review domain. The top shows the training tasks assigned to the 10 clusters. Here the number N$\in\{2,4,5\}$ refers to the threshold of stars for positive reviews. At the bottom we show three tasks with largest improvement from \robusttc-FSL. The top-3 most relevant task clusters (i.e. with highest weights $\alpha$s in Eq.\ref{eqn:fsl} ) are highlighted with \textcolor{blue}{\bf blue bold} font.}
\caption{{\color{red}{\textbf{Precision}}}, {\color{green}{\textbf{Recall}}}, {\color{blue}{\textbf{F-measure}}} and Maximal F-measure ({\color{blue}{$\bullet$}}) of DSS (\textbf{- - -}) and F-DSS (\textbf{---}) under different thresholds. % DSS tends to predict unknown pixels as the majority class--the background, resulting in high precision but low recall. % FLoss is able to find a better compromise between precision and recall. }
\caption{{\color{red}{\textbf{Precision}}}, {\color{green}{\textbf{Recall}}}, {\color{blue}{\textbf{F-measure}}} of model trained under different $\beta^2$ (Eq.~\ref{eq:def-f}). % The precision decreases with the growing of $\beta^2$ whereas recall increases. % This characteristic gives us much flexibility to adjust the balance between recall and precision: use larger $\beta^2$ in a recall-first application and lower $\beta^2$ otherwise. }
\caption{Effect of applying a mapping $f$ to a (disconnected) manifold $\mathscr{M}$ with three hypothetical classes (\textcolor{blue}{\footnotesize{$\blacksquare$}}, \textcolor{red}{$\blacktriangle$} and \textcolor{green}{\large{$\bullet$}}).}
\caption{\label{fig:avAbsPs}Distributions for the absorbing state avalanches (approach I) in three $E$ regimes: (\textcolor{greenC}{$\blacktriangledown$}) subcritical, (\textcolor{redC}{\ding{108}}) critical, and (\textcolor{blueC}{$\blacktriangle$}) supercritical. A: $\Pdist(\sabs)$ and PL fit for the critical regime yielding $\aabs=1.38$. A~inset: Sethna's scaling relation $\sabs\sim\Tabs^a$ yields $a=1.1(1)$ for the critical regime. B: $\Fdist(\sabs)$ and $\Fdist(\Tabs)$ for the three considered regimes. Notice that only the critical regime has PL distributed avalanches.}
\caption{\label{fig:avAbsFsT}Cumulative distributions of absorbing state avalanches (approach I) for the critical regime ($E=1.185\mV$) and four network sizes: (\textcolor{greenC}{$\blacktriangledown$}) $L=20$, (\textcolor{redC}{\ding{108}}) $L=40$, (\textcolor{purpleC}{$\blacksquare$}) $L=80$ and (\textcolor{blueC}{$\blacktriangle$}) $L=99$. A: $\Fdist(\sabs)$. A~inset: plot of $\Hsc(\sabs/L^{\Dabs})\sim\sabs^{\aabs-1}\Fdist(\sabs)$ (collapse of the distributions) yielding $\aabs=1.38$ and $\Dabs=1.75$. B: $\Fdist(\Tabs)$. B~inset: plot of $\Hsc(\Tabs/L^{\muabs})\sim\Tabs^{\tabs-1}\Fdist(\Tabs)$ (collapse of the distributions) yielding $\tabs=1.41$ and $\muabs=0.94$.}
\caption{\label{fig:avSilS}Silent periods avalanche size distributions (approach II). A: $\Pdist(s)$ for $E$ in three regimes: (\textcolor{greenC}{$\blacktriangledown$}) subcritical, (\textcolor{redC}{\ding{108}}) critical, and (\textcolor{blueC}{$\blacktriangle$}) supercritical. Notice that the three regimes have PL shape inside the shaded area. A~inset: critical regime $\Pdist(s)$ for critical $E=1.185\mV$ and different $L$. B: $\Fdist(s)$ for critical $E=1.185\mV$ and different $L$ used to fit Eq.~\eqref{eq:cDistFit} yielding $s_c(L)$ and $\alpha=1.4$. Solid line: example fit of Eq.~\eqref{eq:cDistFit} for $L=20$ (R-Squared: $99.9\%$). B~inset: cutoff $s_c\sim L^D$ yielding $D=0.8(2)$.}
\caption{\label{fig:avSilT}Silent periods avalanche duration distributions (approach II). A: $\Pdist(T)$ for $E$ in three regimes: (\textcolor{greenC}{$\blacktriangledown$}) subcritical, (\textcolor{redC}{\ding{108}}) critical, and (\textcolor{blueC}{$\blacktriangle$}) supercritical. Notice that the three regimes have PL shape inside the shaded area. A~inset: critical regime $\Pdist(T)$ for critical $E=1.185\mV$ and different $L$. B: $\Fdist(T)$ for critical $E=1.185\mV$ and different $L$ used to fit Eq.~\eqref{eq:cDistFit} yielding $T_c(L)$ and $\tau=1.57$. Solid line: example fit of Eq.~\eqref{eq:cDistFit} for $L=20$ (R-Squared: $99.4\%$). B~inset: cutoff $T_c\sim\exp(cL^D)$ yielding $D=1.0(2)$.}
\caption{\label{fig:avPS}Power spectra of avalanche size time series for both approaches and three regimes of $E$: (\textcolor{greenC}{$\blacktriangledown$}) subcritical, (\textcolor{redC}{\ding{108}}) critical, and (\textcolor{blueC}{$\blacktriangle$}) supercritical. A: $\Sps(f)$ for absorbing state avalanches (approach I). The three regimes have white noise power spectrum. B: $\Sps(f)$ for silent periods avalanches (approach II). The critical regime has $\Sps\sim1/f^b$ with $b\sim1.2$.}
\caption{ Variation in enthalpy $h$ as a function of $D/d$ for pressure $p=0.020$. Changing $D/d$ leads to a reversible and continuous transformation between the $(3,2,1)$ uniform structure and the $(3,{\bf 2},1)$ line slip for both the forward (increasing $D/d$, \textcolor{blue}{blue} crosses) and reverse (\textcolor{red}{red} circles) trajectories, indicated by the vertical dashed line. In contrast, the transition from the $(3,{\bf 2},1)$ line-slip structure to $(4,2,2)$ uniform arrangement (thin \textcolor{blue}{blue} line) is discontinuous and occurs at a lower value of $D/d$ on the reverse trajectory (thick \textcolor{red}{red} line). }
\caption{ Variation in enthalpy $h$ as a function of $D/d$ while holding pressure at a constant value of $p=0.026$. The forward trajectory is as before: an increase in $D/d$ leads to a discontinuous transformation from the $(3,2,1)$ uniform structure to the $(3,{\bf 2},1)$ line-slip (thin \textcolor{blue}{blue} line), a further increase in $D/d$ results in a discontinuous transition to the $(4,2,2)$ uniform structure (thin \textcolor{blue}{blue} line). The reverse trajectory is remarkable in that the intervening line-slip is eliminated. Instead the transition from the $(4,2,2)$ uniform structure to the $(3,2,1)$ uniform structure is via a discontinuous transition (thick \textcolor{red}{red} line). The inset shows a zoom on the discontinuous transitions. }
\caption{(Color online) For three globally coupled chaotic R\"ossler oscillators, the influence of the coupling frequency,$\omega$, on network synchronization. (a) The variation of the averaged synchronization error, $\left<\delta x \right>$, with respect to $\omega$ for different coupling amplitudes, $\varepsilon_0=1.60$ and $1.55$. The results are averaged over a time period of $T=1 \times 10^3$ and $1000$ realizations. (b1) The trajectory of the $1$st oscillator in the isolated form in the $(x,y)$ plane. (b2) The PSD of the variable $x$ shown in (b1). The dominant component is locating at $\omega_R\approx 1.1$. (b3) By $\varepsilon_0=1.60$ and $\omega=\omega_2=0.5$, the trajectory of the $1$st oscillator. (b4) By $\varepsilon_0=1.60$ and $\omega=\omega_R$, the trajectory of the $1$st oscillator.} \label{fig1} \end{center} \end{figure*} By the coupling amplitude $\varepsilon_0=1.60$, we plot in Fig. \ref{fig1}(a) the variation of the averaged synchronization error, $\left<\delta x\right>$, with respect to the coupling frequency, $\omega$, for $N=3$ globally coupled R\"{o}ssler oscillators. Here, the averaged synchronization error is defined as $\left<\delta x\right>\equiv\left<\sum_i |x_i-\bar{x}|/N\right>$, with $\bar{x}=\sum_i x_i/N$ the network-averaged state and $\left<\cdots\right>$ the time-average function. Apparently, the smaller is $\left<\delta x\right>$, the stronger is the network synchronized. In addition, to cope with the mismatch of parameters in realistic situations, we set the oscillators with different $\Omega$. Specifically, we set $\Omega=1.0$, $1.0+\delta\omega$ and $1.0-\delta\omega$ for oscillators $1$, $2$ and $3$, respectively, with $\delta\omega=1\times 10^{-3}$. Figure \ref{fig1}(a) shows that, as $\omega$ increases from $0$ to $18$, the value of $\left<\delta x\right>$ is wildly changed. More specifically, when the coupling is static ($\omega=0$), we have $\left<\delta x\right>\approx 4.62$, indicating that the network is deeply desynchronized. However, as $\omega$ increases from $0$, the value of $\left<\delta x\right>$ is quickly decreased and, at about $\omega_1=0.31$, we have $\left<\delta x\right>\approx 0$, indicating that the network work is well synchronized at this point. Increasing $\omega$ further, it is shown that the value of $\left<\delta x\right>$ is gradually increased. After reaching its local maxima at about $\omega_2=0.50$, $\left<\delta x\right>$ begins to decrease again as $\omega$ increases. Then, in a wide range about $\omega\in(0.63,10.08)$, we have $\left<\delta x\right>\approx 0$. This forms the $2$nd window of synchronization in the parameter space of $\omega$. Finally, as $\omega$ exceeds $10.08$, the value of $\left<\delta x\right>$ is gradually increased. Clearly, the network synchronization is modified by tuning $\omega$. It is worth noting that in tuning $\omega$, the amplitude of the coupling strength, $\varepsilon_0$, is kept unchanged. As such, the tuning of $\omega$ provides actually a new approach for synchronization optimization. To check whether the similar phenomenon is observable for other coupling amplitudes, we plot in Fig. \ref{fig1}(a) also the variation of $\left< \delta x \right>$ with respect to $\omega$ for $\varepsilon_0=1.55$. It is seen that in the region of small $\omega$, the variation of $\left<\delta x \right>$ is very close to that of $\varepsilon_0=1.60$. However, in the region of large $\omega$, the value of $\left< \delta x \right>$ is clearly smaller to that of $\varepsilon_0=1.60$. In specific, for $\varepsilon_0=1.55$, the onset of network desynchronization occurs at $\omega\approx13.2$, while for $\varepsilon_0=1.60$ this occurs at $\omega\approx 10.08$. That is, the $2$nd synchronization window is enlarged by decreasing $\varepsilon_0$. Numerical results thus suggest that besides the coupling frequency, the synchronization performance is also influenced by the coupling amplitude. In Ref.~\cite{AB:NODY}, it is reported that when $\omega\approx \omega_R$, the dynamics of the oscillators will be strongly affected by the periodic coupling, resulting in a desynchronization window in the parameter space of $\omega$. For our numerical results shown in Fig. \ref{fig1}(a), there does exist a desynchronization window, $\omega\in (0.48, 0.63)$, yet this window is away from $\omega_R$. On the contrary, as depicted in Fig. \ref{fig1}(a), the network is highly synchronized at $\omega_R$. The contradictory findings intrigue our interest of the impact of $\omega$ on the oscillator dynamics. We first check the dynamics of the oscillators for $\omega=\omega_2$, with which the network is mostly desynchronized. Figure \ref{fig1}(b3) shows the trajectory of the $1$st oscillator in the $(x,y)$ plane (the results for other oscillators are similar). It is seen that the trajectory is of no apparent difference to that of Fig. \ref{fig1}(b1). Meanwhile, the analysis of PSD shows that the dominant frequency is still locating at $\omega_{R}$. We next check the dynamics of the oscillators for $\omega=\omega_R$. The trajectory of the $1$st oscillator is plotted in Fig. \ref{fig1}(b4). Still, the dynamics of the oscillator is not apparently affected by the periodic coupling. Besides the special points $\omega_2$ and $\omega_R$, we have also checked the dynamics of the oscillators for other coupling frequencies (for $\varepsilon_0=1.55$ and $1.60$), and no apparent difference is found between the dynamics of coupled and isolated oscillators. The verification of the oscillator dynamics is necessary and important. On the one hand, it excludes the possibility that the variation of network synchronization, as depicted in Fig. \ref{fig1}(a), is induced by the deformation of the oscillator trajectories. As such, the mechanism revealed in Ref.~\cite{AB:NODY} can not be used to explain the phenomenon observed here. On the other hand, the observation that the trajectories of the oscillators are not (or only slightly) affected by the periodic coupling makes it possible to conduct a theoretical analysis on the numerical results based on the MSF method. In the standard MSF method, the synchronous manifold is of the same dynamics to that of isolated oscillator. It is only under this condition that the curve of MSF is independent of the coupling strength and the network structure~\cite{MSF:Pecora,MSF:HG,MSF:Barahona}. As the calculation of the MSF relies on only the statistical properties of the synchronous manifold (i.e., the largest Lyapunov exponent), the unaffected trajectories under periodic coupling thus suggest that the phenomenon observed in simulations could be analyzed by a unique MSF, as will be detailed in the following section. \section{Mechanism analysis} \subsection{The MSF approach} Although the MSF method is proposed for identical oscillators, it can be applied to oscillators of slight parameter mismatches as well~\cite{SJ:EPL,AS:EPL}. Treating the oscillators as identical, the MSF for periodically coupled oscillators can be obtained, as follows. Let $\bm{x}_s$ be the synchronous manifold of the oscillators and $\delta \bm{x}_i=\bm{x}_i-\bm{x}_s$ be an infinitesimal perturbation added to oscillator $i$, then whether the perturbed trajectories could be converged to the synchronous manifold is mainly determined by the set of variational equations \begin{equation} \delta \dot{\bm{x}}_i=\bm{DF}(\bm{x}_s)\delta \bm{x}_i+\varepsilon(t) \sum_{j=1}^N c_{ij}\bm{DH}(\bm{x}_s)\delta \bm{x}_j, \label{vareq} \end{equation} with $i,j=1,\ldots,N$. Here, $\bm{C}$ is the coupling matrix, with $c_{ij}=a_{ij}$ and $c_{ii}=-k_i=-\sum_j a_{ij}$ for the non-diagonal and diagonal elements, respectively. $\bm{DF}(\bm{x}_s)$ and $\bm{DH}(\bm{x}_s)$ are the Jacobian matrices evaluated on $\bm{x}_s$. For the network to be synchronizable, the necessary condition is that $\delta \bm{x}_i$ decreases to $0$ with time for all the oscillators. Transforming Eqs. (\ref{vareq}) into the mode space spanned by the eigenvectors of the coupling matrix, we have the set of decoupled variational equations \begin{equation} \delta \dot{\bm{y}}_i=\bm{DF}(\bm{x}_s)\delta \bm{y}_i+\varepsilon(t) \lambda_i \bm{DH}(\bm{x}_s)\delta \bm{y}_i, \label{mode} \end{equation} with $\delta\bm{y}_i$ the $i$th mode, and $0=\lambda_1>\lambda_2\geq\lambda_3\geq \ldots\geq \lambda_N$ the eigenvalues of $\bm{C}$. The mode associated with $\lambda_1=0$ characterizes the motion in parallel to the synchronous manifold, whereas the other modes characterize the motion transverse to the synchronous manifold. In the mode space, the necessary condition for synchronization becomes that $\delta\bm{y}_i$ approaches 0 with time for all the transverse modes. Denote $\Lambda_i$ as the largest Lyapunov exponent of Eq. (\ref{mode}), this means that $\Lambda_i$ should be negative for $i=2,\ldots,N$. Introducing the generic coupling strength $\sigma\equiv -\varepsilon_0\lambda$, we have the generalized variational equation \begin{equation} \delta \dot{\bm{y}}=\bm{DF}(\bm{x}_s)\delta \bm{y}- \sigma[1+\sin(\omega t)] \bm{DH}(\bm{x}_s)\delta \bm{y}, \label{MSF} \end{equation} which stands as the MSF of our model. Solving Eq. (\ref{MSF}) numerically, we are able to obtain the variation of $\Lambda$ with respect to $\omega$ and $\sigma$, based on which the stable region of synchronization, i.e., the region with $\Lambda<0$, can be identified. Now, the condition for network synchronization becomes that $\Lambda_i<0$ for $i=2,\ldots,N$, i.e., all the transverse modes should be staying inside of the stable region in the two-dimensional parameter space $(\omega,\sigma)$. \begin{figure*}[tbp] \begin{center} \includegraphics[width=0.95\linewidth]{fig2.eps} \caption{(Color online) The results obtained by the MSF method. (a) The variation of $\Lambda$ with respect to $\omega$. For $\sigma=4.8$ ($\varepsilon_0=1.60$), $\Lambda$ is negative in the regions $\omega\in(0.06,0.44)$ and $\omega\in(0.63,10.08)$. Increasing $\sigma$ to $10$, the stable region is shrunk to $\omega\in(0.63,2.32)$. (b) The dependence of $\Lambda$ on $\sigma$. For static coupling ($\omega=0$), $\Lambda$ is negative in the region $\sigma\in(0.19,4.61)$. By periodic coupling of $\omega=1.27$, the stable region is enlarged to $\sigma\in(0.19, 17.97)$. (c) The contour plot of $\Lambda$ in the parameter space of $(\omega,\sigma)$. The boundaries of the stable region are marked by the dashed lines. (d) The variation of $\omega_r$, i.e., the characteristic frequency for optimized synchronization, with respect to $\sigma$. The fitted data gives $\omega_r\propto \sigma^{-\gamma}$, with $\gamma\approx 0.5$. Squares: the characteristic frequency, $\omega^f_r$, identified by the FLE method. (e) The variation of the network synchronizability, $\beta$, as a function of $\omega$. $\beta$ is diverged at $\omega\approx 0.66$.} \label{fig2} \end{center} \end{figure*} For the network studied in Fig. \ref{fig1}, we have $\lambda_2=\lambda_3=-3$, which, for $\varepsilon_0=1.60$, give $\sigma_2=\sigma_3=\varepsilon_0\lambda_{2,3}=4.8$. Fixing $\sigma=4.8$ in Eq. (\ref{MSF}), we plot in Fig. \ref{fig2}(a) the variation of $\Lambda$ as a function of $\omega$. It is seen that the variation of $\Lambda$ is in good agreement with the numerical results shown in Fig. \ref{fig1}(a). In particular, $\Lambda$ reaches its local minima around $\omega_1=0.31$, and reaches the local maxima around $\omega_2=0.5$ [where $\left<\delta x\right>$ reaches its local maxima in Fig. \ref{fig1}(a)]. Besides confirming the numerical observations, the MSF curve offers more information on the variation of network synchronization. For example, by a closer look at the behavior of $\Lambda$, it is found that the minimum $\Lambda$ is reached at $\omega_r\approx 1.32$, but not at $\omega_R=1.1$. We call $\omega_r$ the characteristic frequency for synchronization. To show the influence of $\sigma$ on MSF, we plot in Fig. \ref{fig2}(a) also the results for $\sigma=10$. Comparing to the case of $\sigma=4.8$, it is seen that the upper bound of the $2$nd synchronization window is significantly decreased (to about $\omega=1.3$). {\color{red}This result is in consistent with the phenomenon observed in Fig. \ref{fig1}(a), in which the $2$nd synchronization window is narrowed as $\varepsilon_0$ is increased from $1.55$ to $1.60$.} Moreover, Fig. \ref{fig2}(a) also shows that for $\sigma=10$, the location of the minimum $\Lambda$, i.e., the characteristic frequency, is shifted slightly to the left ($\omega_r\approx 1.3$). Having justified the validity of the MSF method in quantifying the synchronizability of periodically coupled oscillators, we next employ the this method for a detailed analysis on the dependence of network synchronization on $\omega$ and $\sigma$. Figure \ref{fig2}(b) shows the variation of $\Lambda$ with respect to $\sigma$. For the case of static coupling ($\omega=0$), $\Lambda$ is negative in a bounded region $\sigma\in [0.19, 4.61]$. When periodic coupling with $\omega=1.27$ is adopted, the stable region is enlarged to $\sigma\in [0.19, 17.97]$. Denote $\sigma_l$ and $\sigma_u$ as the lower and upper bounds of the stable region, respectively, the network synchronizability then can be characterized by the ratio $\beta\equiv\sigma_u/\sigma_l$. For the given network structure, the larger is $\beta$, the wider will be the range for synchronization in the parameter space of $\omega$. We thus have $\beta=24.3$ for $\omega=0$ and $\beta=94.58$ for $\omega=1.27$, i.e., the network synchronizability is increased by about four times. Fig. \ref{fig2}(e) shows the detailed variation of $\beta$ with respect to $\omega$. To have a global picture on the dependence of network synchronization on $\omega$ and $\sigma$, we plot in Fig. \ref{fig2}(c) the contour plot of $\Lambda$. Fig. \ref{fig2}(c) shows some interesting features overlooked in previous studies. Firstly, it is shown that by the periodic coupling, the upper bound of the stable region, $\sigma_u$, is significantly affected by varying $\omega$, but the lower bound, $\sigma_l$, is hardly changed. As such, by tuning $\omega$, it is the range of the stable region (or the network synchronizability), $[\sigma_l,\sigma_u]$ (or $\beta$), that is significantly enlarged, whereas the coupling cost (i.e., the smallest coupling amplitude for synchronization) is almost not affected. Secondly, the characteristic frequency where $\Lambda$ reaches its minima, $\omega_r$, in general is different from the oscillator intrinsic frequency, $\omega_R$, and, interestingly, is varying with $\sigma$. As depicted in Fig. \ref{fig2}(d), as $\sigma$ increases, $\omega_r$ is decreased by roughly a power-law scaling. This observation implies that the enhanced synchronization at $\omega_r$ is not induced by the phase-locking between the nodal dynamics and periodic coupling~\cite{AB:NODY}. Thirdly, the upper bound of the stable region is also varied violently at a small coupling frequency (around $\omega_1= 0.31$). This observation confirms again the irrelevance of phase-locking to synchronization enhancement, as this characteristic frequency is also decreased with $\sigma$. Finally, as $\sigma$ increases, the synchronization window is gradually narrowed. Meanwhile, with the increase of $\sigma$, the minimum value of $\Lambda$ is gradually increased. These observations indicate that the enhancement of synchronization by tuning $\omega$ is more prominent for small $\sigma$, which, in realistic situations, corresponds to complex networks of small eigenvalues and weak coupling strength. \subsection{The FLE approach} Whereas the influence of the coupling frequency on synchronization can be analyzed by the MSF method, the underlying mechanism is still not clear. In particular, it remains unknown why $\Lambda$ reaches its minima at $\omega_r$ and why $\omega_r$ is varying with $\sigma$. These questions call for a study on the temporal (local) stability of the MSF. Here we employ FLE for such a purpose~\cite{Pikovsky:Chaos1993,AP:PRE1999,FLE:XGW2006,KS:Chaos2010,AEB:2016}. The FLE is defined as \begin{equation} \Lambda^f(m)=\frac{1}{\Delta T}\ln|\bm{Q}_m(\Delta T)\cdot\bm{u}_0|, \label{FLE} \end{equation} with $\Delta T$ the length of the time interval over which the FLE is averaged, $m$ the interval index, $\bm{Q}({\Delta T})$ the matrix solution of the equation $d\bm{Q}/dt=\bm{DF}(\bm{x}(\Delta T))\cdot\bm{Q}$, and $\bm{u}_0$ the random unit vector in the tangent space of the synchronous manifold. The trajectory is temporally (locally) stable if $\Lambda^f\leq 0$, otherwise it is temporally unstable. For the typical chaotic motion, while the largest Lyapunov exponent is positive, the FLE could be negative for some time intervals. These intervals correspond to regions in the phase space in which infinitesimal vectors in fact contract in length ($\Lambda^f<0$). The asymptotic exponent, i.e., the largest Lyapunov exponent, is just the weighted sum of the temporally positive exponents when the trajectory visits the expanding regions ($\Lambda^f>0$) and the temporally negative exponents when the trajectory is in the contracting regions. For chaotic oscillator, the positive components weight over the negative ones, resulting in the positive asymptotic exponent~\cite{Pikovsky:Chaos1993,AP:PRE1999,FLE:XGW2006}. \begin{figure*}[tbp] \begin{center} \includegraphics[width=0.8\linewidth]{fig3.eps} \caption{(Color online) For $\sigma=8$, the results obtained by the method of FLE analysis. (a) By static coupling, the PSD of FLE. Dashed line denotes the threshold used to identify the dominant components. (b) The variation of the dominant PSD frequencies, $\omega^f$, with respect to the coupling frequency, $\omega$. The vertical dashed lines denote the characteristic frequencies, $\omega^f_1\approx 0.25$ and $\omega^f_r\approx 1$, where the dominant frequencies are broadly distributed. The diagonal line represents the component of the periodic coupling. (c) The PSD of FLE for $\omega=\omega^f_r$. (d) The probability distributions of FLE for $\omega=0$ and $\omega^f_r$.} \label{fig3} \end{center} \end{figure*} We proceed to study the response of FLE of the MSF to periodic coupling. In simulations, we fix $\sigma=8$ and set $\Delta T=0.2$, while noting that the findings to be reported in the following are independent of $\Delta T$ (given that $\Delta T$ is not too large). Fig. \ref{fig3}(a) shows the PSD of FLE for the case of static coupling, $\omega=0$. It is seen that the PSD is featured by the embedding of a few of dominant components over a broad spectral background. To focus on the dominant components, we truncate PSD by the threshold $5\times 10^{-5}$ and record only the dominant frequencies, $\omega^f$. We note that the dominant frequencies characterize the time scales of the local instability of the synchronous manifold, which is different from that of the nodal dynamics. Figure \ref{fig3}(b) shows the variation of the truncated dominant frequencies with respect to $\omega$. Interestingly, it is found that the dominant frequencies are isolated from each other for most values of $\omega$, but are broadly distributed around two characteristic frequencies, $\omega^f_1\approx 0.25$ and $\omega^f_r=1.0$. Remarkably, these characteristic frequencies are very close to the characteristic frequencies, $\omega_{1,r}$, identified by the MSF method [see Fig. \ref{fig2}]. To have more details on the response of FLE at the characteristic frequencies, we plot in Fig. \ref{fig3}(c) the PSD of FLE for $\omega^f_r=1.0$. Comparing with the results of static coupling [Fig. \ref{fig3}(a)], it is seen that in Fig. \ref{fig3}(c) a large number of dominant frequencies are evoked, giving rise to the broad spectral distribution. As a consequence of this, the probability distribution of FLE is changed from the unimodal to bimodal distributions, with the new peak located around $\Lambda^f=-1.52$ [Fig. \ref{fig3}(d)]. It is just the appearance of this new peak that leads to the sharp decrease of $\Lambda$ in the MSF curve. The similar phenomenon is also observed around $\omega^f_1$, where many components are evoked in the PSD of FLE [see Fig. \ref{fig3}(b)] and $\Lambda$ reaches its local mimima in the MSF curve [Fig. \ref{fig2}(a)]. As for the results of MSF analysis, the characteristic frequencies, $\omega^f_{1,r}$, identified by the FLE method are also dependent of the generic coupling strength, $\sigma$. By the same procedure described above, we calculate $\omega^f_r$ for different $\sigma$. The results are also presented in Fig. \ref{fig2}(d). It is seen that $\omega^f_r$ falls exactly on the line fitted by $\omega_r$. The excellent agreement between $\omega^f_r$ and $\omega_r$ manifests the appropriateness of the FLE approach in exploring the synchronization phenomenon of periodically coupled oscillators, and, more importantly, points out the fact that the enhanced synchronization at the characteristic frequencies is due to the resonance between the periodic coupling and the temporal instability of the synchronous manifold (instead of the nodal dynamics). \subsection{Other oscillators} Comparing the FLE spectra of $\omega^f_1$ and $\omega^f_r$ in Fig. \ref{fig3}(b), it is seen that more components are evoked by $\omega^f_r$ than $\omega^f_1$. Meanwhile, in Fig. \ref{fig2}(a) it is shown that comparing to $\omega_1^f$, the value of $\Lambda$ is smaller at $\omega^f_r$. This observation leads to our following hypothesis: the stronger is the resonance between FLE and the periodic coupling, the more significant will be the network synchronization enhanced. We next employ the model of coupled Lorenz oscillators to verify this hypothesis. The Lorenz oscillator in the isolated form is described by equations $(dx/dt, dy/dt, dz/dt )^T = (10y-10x, 28x-xz-y, xy-8z/3)^T$, which presents the chaotic motion in the phase space~\cite{Lorenz}. The oscillators are coupled by the function $\bm{H}([x,y,z]^T)=[0,x,0]^T$, i.e., the $y$ component is coupled to the $x$ component. Still, the coupling strength is varying with time periodically, as described by Eq. (\ref{coupling}). By solving Eq. (\ref{MSF}) numerically, we plot in Fig. \ref{fig4}(a) the contour plot of $\Lambda$ in the two-dimensional parameter space $(\omega,\sigma)$. It is seen that as $\omega$ varies, the range of the stable region, $(\sigma_l,\sigma_u)$, is also modulated. However, comparing to the results of R\"ossler oscillators (Fig.\ref{fig2}), the influence of $\omega$ on synchronization is less prominent. Fixing $\sigma=11$ (where $\sigma_u$ is maximized), we plot in Fig. \ref{fig4}(b) the variation of the dominant frequencies of FLE with respect to $\omega$. Here, $\Delta T=6\times 10^{-2}$ and the truncation threshold is chosen as $2\times 10^{-2}$ (the results for other parameters are qualitatively the same). It is seen that, unlike the case of R\"ossler oscillators[Fig. \ref{fig3}(b)], there is no characteristic frequency by which the distribution of the dominant frequencies is violently changed. \begin{figure}[tbp] \begin{center} \includegraphics[width=0.85\linewidth]{fig4.eps} \caption{(Color online) The results for chaotic Lorenz oscillators. (a) The variation of $\Lambda$ with respect to $\omega$ and $\sigma$. (b) Fixing $\sigma=11$, the variation of the dominant frequencies, as identified from the PSD of FLE, with respect to $\omega$. The diagonal line corresponds to the frequency of periodic coupling.} \label{fig4} \end{center} \end{figure} \section{Discussions and conclusion} The finding that network synchronization is enhanced by periodic coupling provides an alternative approach for synchronization optimization. According to the MSF paradigm, the synchronizability of networked oscillators is jointly determined by two factors: the shape of the MSF curve and the eigenvalues of the network coupling matrix. In conventional studies where constant coupling is adopted, the stable region of the MSF is determined solely by the nodal dynamics and coupling function. As such, to improve the network synchronizability, the only available approach is to modify the eigenvalues of the network coupling matrix. This has led to the extensive studies on the impact of network structure on synchronization over the past two decades~\cite{SYNREV:Boccaletti,SYNREV:Arenas}. The present work, however, adopts a different approach for optimizing synchronization: modifying the MSF curve. By the periodic coupling, the MSF contains two independent parameters, $\sigma$ and $\omega$. As demonstrated in Fig. \ref{fig2}, the stable region of the MSF can be effectively modified by changing $\omega$. In particular, at some characteristic frequencies (e.g., $\omega_r$), the upper bound of the stable region can be significantly increased, making the stable region of MSF significantly enlarged. For network of fixed coupling matrix (i.e., the eigenvalues are fixed), the larger is the stable region, the higher will be the propensity for network synchronization. As a result of the enlarged stable region, the network synchronization is optimized at the characteristic frequencies. A few remarks on the impact of periodic coupling on network synchronization are in order. First, the enlarged stable region of MSF is mainly attributed to the increase of the upper bound, $\sigma_u$. By varying the coupling frequency, the lower bound, $\sigma_l$, is only slightly changed [see Fig. \ref{fig2}]. This restricts the phenomenon of periodic-coupling-enhanced synchronization to be only observable for MSF of bounded stable region. For MSF of semi-open stable region, i.e., only $\sigma_l$ exists, the impact of periodic coupling on synchronization will be less prominent. Second, the enhanced synchronization at the characteristic frequencies, $\omega_{1,r}$, is rooted in the nonlinear response of the temporal instability of the synchronous manifold, i.e., the FLE, to the periodic driving. To realize the nonlinear response, a necessary condition is that the coupling matrix must be non-diagonal, i.e., $\bm{H}([x,y,z]^T)\neq [x,y,z]^T$. If the coupling matrix is diagonal, the stable region will be not affected by the periodic coupling, even when the time scales of the periodic coupling and the nodal dynamics are comparable. Third, whereas our study points out the connection between the periodic coupling and the temporal stability of the synchronous manifold, further studies are still needed to understand the mechanism of enhanced synchronization by periodic coupling. As a simple measure of the temporal stability, we employ FLE in the present work to investigate the response of MSF to periodic coupling. Although violent changes are observed in FLE at the characteristic frequencies, it remains unknown to us how this happens. It is our hope that this question could be addressed in the near future by examining more examples and employing new mathematical techniques. To summarize, we have studied the synchronization behavior of networked chaotic oscillators with periodic coupling, and found that the network synchronization can be effectively enhanced by tuning the coupling frequency. Based on the method of MSF, we have conducted a detailed analysis on the impact of the coupling frequency on synchronization, and found that network synchronization is optimized at some characteristic frequencies comparable to that of the nodal dynamics. The mechanism of synchronization optimization is investigated by the technique of FLE, and it is revealed that the optimized synchronization is due to a violent change of the PSD of FLE at the characteristic frequencies. Our study sheds new lights on the synchronization behavior of temporally coupled oscillators, and the findings might have applications to the optimization of synchronization in realistic networks. This work was supported by the National Natural Science Foundation of China under the Grant Nos.~11375109 and 61703257, and by the Fundamental Research Funds for the Central Universities under the Grant No.~GK201601001. \begin{thebibliography}{99} %Review \bibitem{SYNBOOK:Kuramoto} Y. Kuramoto, \emph{Chemical Oscillations, Waves, and Turbulence} (Springer, Berlin, 1984). \bibitem{SYNBOOK:Pikovsky} A. Pikovsky, M. Rosenblum, and J. Kurths, \emph{Synchronization: A Universal Concept in Nonlinear Science} (Cambridge University Press, Cambridge, 2003). \bibitem{SYNBOOK:Strogatz} S. H. Strogatz, \emph{Sync: The Emerging Science of Spontaneous Order} (Hyperion, New York, 2003). \bibitem{SYNREV:Boccaletti} S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, and D.-U. Hwang, Complex networks: Structure and dynamics, Phys. Rep. {\bf 424}, 175 (2006). \bibitem{SYNREV:Arenas} A. Arenas, A. D\'iaz-Guilera, J. Kurths, Y. Moreno, and C. S. Zhou, Synchronization in complex networks, Phys. Rep.{\bf 469}, 93 (2008). %MSF \bibitem{MSF:Pecora} L. M. Pecora and T. L. Carroll, Master stability functions for synchronized coupled systems, Phys. Rev. Lett. {\bf 80}, 2109 (1998). \bibitem{MSF:HG} G. Hu, J. Z. Yang, and W. J. Liu, Instability and controllability of linearly coupled oscillators: Eigenvalue analysis, Phys. Rev. E {\bf 58}, 4440 (1998). \bibitem{MSF:Barahona} M. Barahona and L. M. Pecora, Synchronization in small-world systems, Phys. Rev. Lett. {\bf 89}, 054101 (2002). % Net Syn \bibitem{NETSYN:CKHU} P. M. Gade and C.-K. Hu, Synchronous chaos in coupled map lattices with small-world interactions, Phys. Rev. E {\bf 62}, 6409 (2000). \bibitem{NETSYN:Nishikawa} T. Nishikawa, A. E. Motter, Y.-C. Lai, and F. C. Hoppensteadt, Heterogeneity in oscillator networks: Are smaller worlds easier to synchronize? Phys. Rev. Lett. {\bf 91}, 014101 (2003). \bibitem{NETSYN:Motter2005} A. E. Motter, C. Zhou, and J. Kurths, Weighted networks are more synchronizable: How and why, AIP Conf. Proc. {\bf 776}, 201 (2005). \bibitem{NETSYN:WXG2007} X. G. Wang,Y.-C.Lai, and C.-H. Lai, Enhancing synchronization based on complex gradient networks, Phys. Rev. E {\bf 75}, 056205 (2007). %temporal network \bibitem{TemNet:Holme} P. Holme and J. Saram\"aki, Temporal networks, Phys. Rep.{\bf 519}, 97 (2012). \bibitem{TemNet:LiAming} A. Li, S. P. Cornelius, Y.-Y. Liu, L. Wang, and A.-L. Barab\'asi, The fundamental advantages of temporal networks, Science\textbf{358}, 1042 (2017). %evolutionary network \bibitem{SND:2002} S. N. Dorogovtsev and J. F. F. Mendes, Evolution of networks, Adv. Phys. \textbf{51}, 1079 (2002). \bibitem{SG:2007}G. Szab\'o and G. F\'ath, Evolutionary games on graphs, Phys. Rep.{\bf 446}, 97 (2007). \bibitem{ZCS:2006} C. Zhou, J. Kurths, Dynamical weights and enhanced synchronization in adaptive complex networks, Phys. Rev. Lett. \textbf{96}, 164102 (2006). \bibitem{FW:2011} C. Fu and X. G. Wang, Network growth under the constraint of synchronization stability, Phys. Rev. E \textbf{83}, 066101 (2011). \bibitem{MWFDL:2011} M. Li, X. G. Wang, Y. Fan, Z. Di, and C.-H. Lai, Onset of synchronization in weighted complex networks: The effect of weight-degree correlation, Chaos \textbf{21}, 025108 (2011). \bibitem{MOTTER:2005} A. E. Motter and Y.-C. Lai, Cascade-based attacks on complex networks, Phys. Rev. E {\bf 66}, 065102 (2002). \bibitem{Holme:2006} P. Holme and M. E. J. Newman, Nonequilibrium phase transition in the coevolution of networks and opinions, Phys. Rev. E \textbf{74}, 056108 (2006). \bibitem{IBS:2010} I. B. Schwartz and L. B. Shaw, Rewiring for adaptation, Physics {\bf 3}, 17 (2010). \bibitem{WYF:2016} Y. F. Wang, H. W. Fan, W. J. Lin, Y.-C. Lai, and X. G. Wang, Growth, collapse, and self-organized criticality in complex networks, Sci. Rep. {\bf 6}, 24445 (2016). %time-dependent coupling \bibitem{Belykh:2004} I. V. Belykh, V. N. Belykh, and M. Hasler, Blinking model and synchronization in small-world networks with a time-varying coupling, Physica D {\bf 195}, 188 (2004). \bibitem{Boccaletti:2006} S. Boccaletti, D.-U. Hwang, M. Chavez, A. Amann, J. Kurths, and L. M. Pecora, Synchronization in dynamical networks: Evolution along commutative graphs, Phys. Rev. E {\bf 74}, 016102 (2006). \bibitem{CM:2007} M. Chen, Synchronization in time-varying networks: A matrix measure approach, Phys. Rev. E {\bf 76}, 016104 (2007). \bibitem{SO:2008}F. Sorrentino and E. Ott, Adaptive synchronization of dynamics on evolving complex networks, Phys. Rev. Lett. \textbf{100}, 114101 (2008). \bibitem{FM:2008PRL} M. Frasca, A. Buscarino, A. Rizzo, L. Fortuna, and S. Boccaletti, Synchronization of moving chaotic agents, Phys. Rev. Lett. {\bf 100}, 044102 (2008). \bibitem{CL:2009PRE} L. Chen, C. Qiu, and H. B. Huang, Synchronization with on-off coupling: Role of time scales in network dynamics, Phys. Rev. E {\bf 79}, 045101 (2009). \bibitem{CL:2010EPJB} L. Chen, C. Qiu, H. B. Huang, G. X. Qi, and H. J. Wang, Facilitated synchronization of complex networks through a discontinuous coupling strategy, Eur. Phys. J. B {\bf 76}, 625 (2010). \bibitem{Porfiri:2012} M. Porfiri, Stochastic synchronization in blinking networks of chaotic maps, Phys. Rev. E {\bf 85}, 056114 (2012). \bibitem{Hasler:SIADS2} M. Hasler, V. N. Belykh, and I. V. Belykh, Dynamics of stochastically blinking systems, Part I: Finite time properties, SIAM J. Appl. Dyn. Syst. {\bf 12}, 1007 (2013). \bibitem{VK:2014} V. Kohar, P. Ji, A. Choudhary, S. Sinha and J. Kurths, Synchronization in time-varying networks, Phys. Rev. E {\bf 90}, 022812 (2014). \bibitem{TIMME:PRL} M. Schr\"oder, M. Mannattil, D. Dutta, S. Chakraborty and M. Timme, Transient uncoupling induces synchronization, Phys. Rev. Lett.{\bf 115}, 054101 (2015). \bibitem{ZhouJ:2016} J. Zhou, Y. Zou, S. G. Guan, Z. H. Liu, and S. Boccaletti, Synchronization in slowly switching networks of coupled oscillators, Sci. Rep. {\bf 6}, 35979 (2016). \bibitem{FN:2016Chaos} N. Fujiwara, J. Kurths, and A. Di\'az-Guilera, Synchronization of mobile chaotic oscillator networks, Chaos{\bf 26}, 094824 (2016). \bibitem{AB:NODY} A. Buscarino, M. Frasca, M. Branciforte, L. Fortuna, J. C. Sprott, Synchronization of two R\"ossler systems with switching coupling, Nonlinear Dyn.{\bf 88}, 673 (2017). \bibitem{Golovneva:2017} O. Golovneva, R. Jeter, I. Belykh, M. Porfiri, Windows of opportunity for synchronization in stochastically coupled maps, Physica D {\bf 340}, 1 (2017). %others \bibitem{Rossler} O. E. R\"ossler, An equation for continuous chaos, Phys. Lett. A{\bf 57}, 397 (1976). \bibitem{SJ:EPL} J. Sun, E. M. Bollt, and T. Nishikawa, Master stability functions for coupled nearly identical dynamical systems, EPL {\bf 85}, 60011 (2009). \bibitem{AS:EPL} S. Acharyya and R. E. Amritkar, Synchronization of coupled nonidentical dynamical systems, EPL {\bf 99}, 40005 (2012). %FLE \bibitem{Pikovsky:Chaos1993} A. Pikovsky, Local Lyapunov exponents for spatiotemporal chaos, Chaos {\bf 3}, 225 (1993). \bibitem{AP:PRE1999} A. Prasad and R. Ramaswamy, Characteristic distributions of finite-time Lyapunov exponents, Phys. Rev. E {\bf 60}, 2761 (1999). \bibitem{FLE:XGW2006} X. G. Wang, Y.-C. Lai, and C.-H. Lai, Characterization of noise-induced strange nonchaotic attractors, Phys. Rev. E {\bf 74}, 016203 (2006). \bibitem{KS:Chaos2010} K. Stefa\'nski, K. Buszko, and K. Piecyk, Transient chaos measurements using finite-time Lyapunov exponents, Chaos{\bf 20}, 033117 (2010). \bibitem{AEB:2016} A. E. Botha, Characteristic distribution of finite-time Lyapunov exponents for chimera states, Sci. Rep. {\bf 6}, 29213 (2016). \bibitem{Lorenz} E. N. Lorenz, Deterministic nonperiodic flow, J. Atmos. Sci. {\bf 20}, 130 (1963). \end{thebibliography} \end{document} \bibitem{WS_model:Watts} D. J. Watts and D. H. Strogatz, Collective dynamics of `small-world' network, Nature (London) {\bf 393}, 440 (1998). \bibitem{NW_model:Newman} M. E. J. Newman and D. J. Watts, Renormalization group analysis of the small-world network model, Phys. Lett. A {\bf 263}, 341 (1999). \bibitem{ER_model:Erdos} P. Erd\"os and A. R\'enyi, On the evolution of random graphs, Publ. Math. Inst. Hung. Acad. Sci.{bf 5},17 (1960). \bibitem{Chen_system:Chen} G. Chen and T. Ueta, Yet another chaotic attractor, Int. J. Bifurcation Chaos Appl. Sci. Eng. {\bf 9}, 1465(1999). \bibitem{HR_system:Dhamala} M. Dhamala, V. K. Jirsa and M. Ding, Transitions to Synchrony in Coupled Bursting Neurons, Phys. Rev. Lett. {\bf 92}, 028101 (2004). \bibitem{Duffing_system:Stefanski} A. Stefa\'nski, P. Perlikowski and T. Kapitaniak, Ragged synchronizability of coupled oscillators, Phys. Rev. E{\bf 75}, 016210 (2007). \bibitem{Pol_system:Mettin} R. Mettin, U. Parlitz, and W. Lauterborn, Bifurcation structure of the driven van der Pol oscillator, Int. J. Bifurcation Chaos Appl. Sci. Eng. {\bf 3}, 1529 (1993). }
\caption{\label{figure:results}For each dataset, the first row is the adjusted rand index scores and the second row is the adjusted mutual information scores. Bolded are {\color{OliveGreen} \bf highest} and {\color{orange} \bf second highest} scores. For MCores and Quickshift++, we used a single $\beta = 0.3$ for each dataset with the exception of for banknote where $\beta = 0.7$. Then the procedures were tuned in their respective essential hyperparameter: $k$-means (KMns) number of clusters, DBSCAN (DScn) epsilon, MCores (MCrs) $k$ from $k$-NN, mean shift (MSft) bandwidth, quick shift (QSft) bandwidth, Quickshift++ (QS++) $k$.}
\caption{ETH3D Benchmark evaluation in the high-resolution multi-view scenario (\textcolor[rgb]{0.5,0.5,0.5}{indoor} and \textcolor[rgb]{0.25,0.875,0.8125}{outdoor} datasets) for different methods, including \textcolor[rgb]{1,0,0}{PMVS}~\cite{Alpher01}, \textcolor[rgb]{1,0.84,0}{CMPMVS}~\cite{Alpher20}, \textcolor[rgb]{0,0,1}{Gipuma}~\cite{Alpher09}, \textcolor[rgb]{0,1,0}{COLMAP}~\cite{Alpher10} and \textcolor[rgb]{0,0,0}{ours (AMHMVS)}. Results are shown as a solid line for accuracy and as a dashed line for completeness. The related values are from ~\cite{Alpher34} (Best viewed in color).}
\caption{ Correlation function of the Langevin random-force. Continuous curve refers to the normalized Planck black-body law of Eq. (\ref{cf_planck}) and dashed curve refers to the normalized\colorbox[rgb]{1,0,0}{ Gaussian shape $1/\sqrt{\pi} exp(t/\tau_{\epsilon})$}. Both curves look significantly broadened with respect to the classical delta-function shape, with the Planck curve exhibiting a longer tail than the Gaussian curve. \label{fig6} }
\caption{ Human analysis of the context required to answer questions on TriviaQA (Wikipedia). 50 examples are sampled randomly. `N sent' indicates the number of sentences required to answer the question, and `N/A' indicates the question is not answerable even given all sentences in the document. The groundtruth answer text is in \red{red text}. Note that the span is not given as the groundtruth. In the first example classified into `N/A', the question is not answerable even given whole documents, because there is no word `corlourful' or `enchantment' in the given documents. In the next example, the question is also not answerable even given whole documents, because all sentences containing `London' does not contain any information about Canaletto's landscapes. }
\caption{Examples on SQuAD, which \ours~predicts the wrong answer. Grountruth span is in {\underline{underlined text}}, the prediction from \ours~is in \red{red text}. Sentences selected by \ourselector~is denoted with \red{\checkmark}. In the first example, the model predicts the wrong answer from the oracle sentence. In the second example, the model predicts the answer from the wrong sentence, although it selects the oracle sentence. In the last example, the model fails to select the oracle sentence. }
\caption{Examples on SQuAD-Adversarial. Groundtruth span is in {\underline{underlined text}}, and predictions from \full~and \ours~are in \blue{blue text} and \red{red text}, respectively.}
\caption{ Human analysis of the context required to answer questions on SQuAD and TriviaQA. 50 examples from each dataset are sampled randomly. `N sent' indicates the number of sentences required to answer the question, and `N/A' indicates the question is not answerable even given all sentences in the document. `Document' and `Question' are from the representative example from each category on SQuAD. Examples on TriviaQA are shown in Appendix~\ref{sec:app-analysis}. The groundtruth answer span is in \red{red text}, and the oracle sentence (the sentence containing the grountruth answer span) is in {\bf bold text}. }
\caption{ Examples on SQuAD. Grountruth span ({\underline{underlined text}}), the prediction from \full~(\blue{blue text}) and \ours~(\red{red text}). Sentences selected by \ourselector~is denoted with \red{\checkmark}. In the above two examples, \ours~correctly answer the question by selecting the oracle sentence. In the last example, \ours~fails to answer the question, since the inference over first and second sentences is required to answer the question.}
\caption{An example on SQuAD, where the sentences are ordered by the score from \ourselector. Grountruth span ({\underline{underlined text}}), the predictions from \topone~(\blue{blue text}), \toptwo~(\green{green text}) and \dyn~(\red{red text}). Sentences selected by \topone, \toptwo~and \dyn~are denoted with \blue{\checkmark}, \green{\checkmark} and \red{\checkmark}, respectively. %By selecting two sentences, \dyn~is able to choose the oracle sentence. (Top) \topone~fails to select the oracle sentence, hence predicts the wrong answer. (Bottom) \toptwo~predicts the answer from the wrong sentence. }
\caption{Attention maps for $\softmax$ and $\csparsemax$ for two {\sc De-En} sentence pairs (white means zero attention). Repeated words are {\color{red}{highlighted}}. The reference translations are \emph{``This is Moore's law over the last hundred years''} and \emph{``I am going to go ahead and select government.''}}
\caption{Example of cross-lingual information extraction: Chinese input text (a) and linearized English PredPatt output (b), where `:p' and \textcolor{blue}{blue} stand for predicate while `:a' and \textcolor{purple}{purple} denote argument.\label{fig:repr}}
\caption{Visualization of {\em Halo} method. While a neural model learns to summarizes the current known information into a hidden state and predict the next target token, the surroundings of this hidden state in the same space (two-dimensional in this example) are supervised to generate tokens with the same semantic structure tag. For example, at the last shown step, the centroid of \textcolor{purple}{purple} area is the summarized hidden state and learns to predict `mortars:a', while a randomly sampled neighbor is enforced to generate an argument, although it may not be `mortars' (thus denoted by `?'). Similar remarks apply to the \textcolor{blue}{blue} regions.}
\caption{(Color online) (a) Topological indices $\mathcal{O}_{Z^{~}_{4} \times Z^{~}_{4}}$ and $\mathcal{O}_\mathcal{I}$, and (b) expectation values of the quadratic bond Casimir operator $\langle ({\bm T}_j +{\bm T}_{j+1})^2 \rangle$ as a function of the coupling $J/J^{~}_{\rm g}$ given in the extended SU(4) Hamiltonian~(\ref{eq:hamiltonian_eg_hb}). The case $J=0$ corresponds to the SU(4) $e$-$g$ spin model and $J/J_g=1$ to the SU(4) bilinear model. The symbols $\bigcirc$, $\bullet$, and \textcolor{red}{$\times$} represent the indexes $\mathcal{O}_{Z^{~}_{4} \times Z^{~}_{4}}$ of $A^{\sigma}$, $\mathcal{O}_{Z^{~}_{4} \times Z^{~}_{4}}$ of $(A^{\sigma})^{T}$, and $\mathcal{O}_\mathcal{I}$, respectively. The auxiliary space of iDMRG calculations is set as $\chi=225$. }
\caption{Numerical proof of the existence of four types of minimizers for the density $f_2$ of Section \ref{sec:code}. Colors correspond to types as: \textcolor{violet}{purple} = 213; \textcolor{blue}{blue} = 3123; \textcolor{teal}{green} = 2313; \textcolor{orange}{yellow} = 32123. See more information in Figure \ref{fig:seconddensity}.}
\caption{Snapshots from the animation (\url{https://github.com/natso26/triple-bubbles/raw/master/triple.mp4}) of 3D triple bubble type as a function of the three prescribed volumes $V_1$, $V_2$, $V_3$ for the density $f_1$ \eqref{eq:density1}. Note that the $V_2$ and $V_3$ axes are on different scales. Colors correspond to types as: \textcolor{violet}{purple} = 213; \textcolor{blue}{blue} = 3123; \textcolor{orange}{yellow} = 32123. The white areas near the axes are regions where $V_1 \leq V_2 \leq V_3$ is not satisfied and so nothing is drawn. As $V_1$ increases, the blue region pushes the yellow region to the right, while the purple region slowly rises.}
\caption{Snapshots from the animation (\url{https://github.com/natso26/triple-bubbles/raw/master/triple2.mp4}) of 3D triple bubble type as a function of the three prescribed volumes $V_1$, $V_2$, $V_3$ for the density $f_2$ \eqref{eq:density2}. The $V_2$ and $V_3$ axes are on different scales. Colors correspond to types as: \textcolor{violet}{purple} = 213; \textcolor{blue}{blue} = 3123; \textcolor{teal}{green} = 2313; \textcolor{orange}{yellow} = 32123. The white areas near the axes are regions where $V_1 \leq V_2 \leq V_3$ is not satisfied and so nothing is drawn. The white stripes near the transition boundaries are regions where the difference between a perimeter-minimizing type and another type does not exceed $10^{-4}$ and so nothing is drawn to guard against rounding errors. \captionpar\setlength{\parindent}{1em} In this animation, $V_1$ increases exponentially. As $V_1$ increases, the purple region first squeezes the green region out of existence; then the blue region emerges and pushes the yellow region to the right while the purple region slowly rises. The green region can only be seen when $V_1$ is small (less than 0.05).}
\caption{SDRNN architecture for sequence processing, unrolled in time such that each column denotes a single time step with a corresponding \textcolor{nnin}{input} and \textcolor{nnout}{output}. The \textcolor{nnattin}{hidden state} is denoised by the \textcolor{nnattrnet}{attractor net}, yielding a \textcolor{nnattout}{cleaned state} which is combined with the next \textcolor{nnin}{input} to determine the next \textcolor{nnattin}{hidden state}.}
\caption{Visual tracking comparison between \textcolor{red}{\textbf{PBTS}} and the best three part-based trackers on the VOT2018 benchmark, \textcolor{ForestGreen}{ANT}, \textcolor{blue}{DPT}, and \textcolor{Goldenrod}{LGT}, on the sequences \textit{motocross1}, \textit{ants1}, and \textit{zebrafish1} (first, second, and third rows respectively). Frames in which there is no bounding box for a tracker indicate that it failed recently and will be reinitialised.}
\caption{Failure cases in challenging VOT2018 benchmark sequences (\textit{bolt1}, \textit{rabbit}, and \textit{gymnastics2}). Tracked object parts are shown in \textcolor{red}{red}, with the predicted and ground-truth bounding boxes shown in \textcolor{ForestGreen}{green} and \textcolor{cyan}{cyan} respectively.}
\caption{\textbf{A snapshot of our web based image annotation system}. Rather than starting from scratch, an annotator is provided with five tags and five sentences automatically recommended by the system based on the pictorial content. When clicked, the tags and sentences will instantly appear in the editable text forms to assist manual annotation. \blue{Source code of the annotation system is publicly available at https://tinyurl.com/cococn-system}.}
\caption{ Time evolution of $\langle x^2\rangle_c$, $\langle x^2\rangle_s$ and $\langle y^2\rangle_c$, $\langle y^2\rangle_s$ in the case of a pure deformation. Parameters are similar to that of experiments ($D_s=1360\,\mu\mathrm{m^2s^{-1}}$, $D_c=2\,\mu\mathrm{m}^2s^{-1}$, $D_\mathrm{dp}=290\,\mu\mathrm{m^2\, s^{-1}}$), $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$, and $\sigma=1\,\mathrm{s}^{-1}$, corresponding to $Pe_c=5.\,10^5$ and $Pe_s=735$. \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{magenta}{\textbf{$-\!\cdot\!\cdot\!-\!\cdot\!\cdot\!-$}}: salt. }
\caption{ Time evolution of $\langle X'^2\rangle_c$, $\langle X^2\rangle_s$ and $\langle Y'^2\rangle_c$, $\langle Y^2\rangle_s$ for $\gamma=1\,\mathrm{s}^{-1}$. Parameters are the same as in the experiments ($D_s=1360\,\mu\mathrm{m^2s^{-1}}$, $D_c=2\mu\mathrm{m}^2s^{-1}$ and $D_\mathrm{dp}=290\mu\mathrm{m^2s^{-1}}$), $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$, corresponding to $Pe_c=5.\,10^5$ and $Pe_s=735$. \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{magenta}{\textbf{$-\!\cdot\!\cdot\!-\!\cdot\!\cdot\!-$}}: evolution for the salt. }
\caption{ Time evolution of $\tilde c=\sqrt{\Delta_c(0)/\Delta_c}$ in the case of pure deformation (left) and pure shear (right). Same set of parameters as in experiments ($D_s=1360\,\mu\mathrm{m^2s^{-1}}$ and $D_c=2\,\mu\mathrm{m}^2s^{-1}$, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$), $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$. Pure deformation case, $\sigma=1\,\mathrm{s}^{-1}$, $Pe_c=5.\,10^5$. Pure shear $\gamma=1\,\mathrm{s}^{-1}$, $Pe_c=5.\,10^5$. Legends are \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{magenta}{\textbf{$-\!\cdot\!\cdot\!-\!\cdot\!\cdot\!-$}}: case of salt. The horizontal line $y=1/2$ serves as measuring the mixing time, $T_\mathrm{mix}$, such that $\tilde c (T_\mathrm{mix}) = 1/2$. }
\caption{ Evolution of the mixing time $T_\mathrm{mix}$ in the case of pure deformation (left, $\sigma=1\,\mathrm{s}^{-1}$) and pure shear (right, $\gamma=1\,\mathrm{s}^{-1}$) for a fixed shear rate and variable colloid diffusion coefficient $D_c \in [0.2, 2 \cdot 10^{5}]\mu\mathrm{m}^2s^{-1}$. Same set of parameters as in experiments $D_s=1360\,\mu\mathrm{m^2s^{-1}}$, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$, $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$. Legends are \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$. }
\caption{ Time evolution of $\tilde c=\sqrt{\Delta_c(0)/\Delta_c}$ in the case of pure deformation (left, $\sigma=1\,\mathrm{s}^{-1}$) and pure shear (right, $\gamma=1\,\mathrm{s}^{-1}$) for a fixed shear rate and four values of the colloid diffusion coefficient $D_c=[0.02, 2, 200, 20000]\mu\mathrm{m}^2s^{-1}$. Same set of parameters as in experiments $D_s=1360\,\mu\mathrm{m^2s^{-1}}$, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$, $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$. Legends are \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$. }
\caption{ Time evolution of $\tilde c=\sqrt{\Delta_c(0)/\Delta_c}$ for a fixed colloid diffusion coefficient and different values of the shear. Left, pure deformation $\sigma=1\,\mathrm{s}^{-1}$, $0.5\,\mathrm{s}^{-1}$, $0.25\,\mathrm{s}^{-1}$, $0.125\,\mathrm{s}^{-1}$; right, pure shear $\gamma=0.02\,\mathrm{s}^{-1}$, $0.1\,\mathrm{s}^{-1}$, $1\,\mathrm{s}^{-1}$, $10\,\mathrm{s}^{-1}$. Same set of parameters as in experiments ($D_s=1360\,\mu\mathrm{m^2s^{-1}}$, $D_c=2\mu\mathrm{m}^2s^{-1}$, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$), $y_{0,c}^2=x_{0,c}^2=1\,\mathrm{mm^2}$. Legends are \textcolor{red}{\textbf{$-\cdot-\cdot-$}}: ``salt-attracting" case, $D_\mathrm{dp}=290\,\mu\mathrm{m}^2s^{-1}$; \textcolor{black}{---}: ``no-salt" case, $D_\mathrm{dp}=0\,\mu\mathrm{m^2\, s^{-1}}$; \textcolor{blue}{- - -}: ``salt-repelling" case, $D_\mathrm{dp}=-290\,\mu\mathrm{m^2\, s^{-1}}$. }
\caption{Evolution of the relative maximum measured from $\tilde c(t)$ as a function of the predicted value using equation (\ref{eq:c_tilde_max2}). Legends are (\textcolor{blue}{\textbf{$\circ$}}): pure shear. (\textcolor{red}{\textbf{$\square$}}): pure deformation. (\textcolor{black}{-~-}): straight line of equation $y=x$.}
\caption{\footnotesize Attribute-based Maximum Precision Rate and Maximum Success Rate (MPR/MSR \%) of different tracking algorithms on RGBT234 dataset, including ECO~\cite{danelljan2017eco}, C-COT~\cite{danelljan2016beyond}, CFnet~\cite{Cfnet2017cvpr}, CSR-DCF~\cite{lukezic2017csrdcf}, SRDCF~\cite{danelljan2015learning}, SAMF~\cite{li2014scale}, DSST~\cite{danelljan2014accurate}, SOWP~\cite{Kim15iccv}, KCF~\cite{CSK15pami}, MEEM~\cite{MEEM14eccv}, CSR~\cite{Li16tip}, JSR~\cite{Liu12infosci} and L1-PF~\cite{Wu11icif}. The best, second and third best performance are marked in \textcolor{red}{red}, \textcolor{green}{green} and \textcolor{blue}{blue} colors, respectively.}
\caption{Linear and log-plots of $\p(\yxs = \yrs)$ constructed by \me\at the population level followed by sample marginalization (blue triangles),\eqn~\eqref{eq:app_maxent_pop_marg}, and at the sample level (red circles), \eqn~\eqref{eq:app_maxent_sample}, with $\yNv=5000$, $n=200$, and constraints as in \eqn~\eqref{eq:constraints}.}
\caption{Linear and log-plots of $\p(\yxs = \yrs)$ constructed by \me\at the population level followed by sample marginalization (blue triangles),\eqn~\eqref{eq:app_maxent_pop_marg}, and at the sample level (red circles), \eqn~\eqref{eq:app_maxent_sample}, with $\yNv=5000$, $n=200$, and constraints as in \eqn~\eqref{eq:constraints}.}
\caption{Linear and log-plots of $\p(\yxs = \yrs)$ for a sample of $n=200$ and constraints as in \eqn~\eqref{eq:constraints}, constructed by: \textcolor{myred}{\textbf{red squares:}} \me\at the sample level,\eqn~\eqref{eq:app_maxent_sample}; \textcolor{mybluishpurple}{\textbf{blue triangles:}} \me\at the population level,\eqn~\eqref{eq:app_maxent_pop_marg} with $\yNv=10\,000$, followed by sample marginalization; \textcolor{myyellow}{\textbf{yellow circles:}} \me\at the population level with unknown population size,\eqn~\eqref{eq:app_maxent_pop_marg_unknown_N}, according to the distribution~\eqref{eq:pdf_popsize} for the population.}
\caption{A schematic representation of the IEEE 33-bus distribution test system \cite{33_bus}, where the root bus is denoted in green (\#~1), bus with distributed generator in orange (\#~14), buses with PV resources in yellow (\#~2, 3, 6, 18, 21, 25, 32), and buses with TCL ensembles in blue (\#~17, 20, 23, 26).}
\caption{ {\bf Average X-ray luminosity of the least massive accretor versus the ratio of the inner binary's final mass to its initial mass (upper panel) and final orbital period (lower panel).} As in the previous figures, systems in: {\color{Cyan}} transferred mass entirely through winds; {\color{red}} transferred mass entirely through winds and then produced a common envelope when the donor filled its Roche lobe; {\color{Blue}} transferred mass through winds and then through the L1 point as well, during Roche-lobe filling. The luminosity of $M_2$ is computed in each time step and ts average value over the time during which it is larger than $10^{32}$~erg~s$^{-1}$ is computed. }
\caption{Performance of our policy gradient method against \redqueen{}~\cite{redqueen17wsdm} and Karimi's method~\cite{karimi2016smart} on feeds sorted in reverse chronological order. % Panels (a) and (b) show the average rank and time at the top, where the solid horizontal line shows the median value across users, normalized with respect to the value achieved by a user who follows a uniform Poisson intensity, and the box limits correspond to the $25$\%-$75$\% percentiles. For the average rank, lower is better and, for time at the top, higher is better. In both cases, the number of messages posted by each method is the same. % % Panel (c) shows a user'{}s intensity $\uStar(\cdot)$ (in blue), as provided by our method, the counts of the user'{}s posts (in green), and the average % rank (in red). }
\caption{Smart broadcasting. Performance of our policy gradient method against \redqueen{}~\cite{redqueen17wsdm} (RQ), a variant of \redqueen{} which has access to true ranks (RQ$^*$), and Karimi's method~\cite{karimi2016smart} on feeds using a sorting algorithm based on a priority queue (refer to Appendix~\ref{app:feed-sorting-algorithm}). % Panels (a) and (b) show the average rank and time at the top, where the solid horizontal line shows the median value across users, normalized with respect to the value achieved by a user who follows a uniform Poisson intensity, and the box limits correspond to the $25$\%-$75$\% percentiles. For the average rank, lower is better and, for time at the top, higher is better. In both cases, the number of messages posted by each method is the same. % Panel (c) shows a user'{}s intensity $\uStar(\cdot)$ (in blue), as provided by our method, the counts of the user'{}s posts (in green), the average rank (in red), the posting times of a competing user with higher priority (in purple), and the posting times of another competing user with lower priority (in yellow). }
\caption{ Example predictions by \modelname~on the development split of the WoZ restaurant reservation dataset. Model predicted slot-value pairs that are not in the ground truth (e.g. {\color{blue} false positives}) are prefaced with a ``{\color{blue} +}'' symbol. Ground truth slot-value pairs that are not predicted by the model (e.g. {\color{red} false negatives}) are prefaced with a ``{\color{red} -}'' symbol. }
\caption{Number of dendrogram features as a function of intensity spacing, $\delta$, for all models. The lines indicate power-law fits. \textcolor{red}{}}
\caption{Comparing the results of OWSC, OWTC and ScaleSSKCF based on mean distance precision (DP) and mean overlap precision (OP). The entries in \textcolor[rgb]{1.00,0.00,0.00}{red} denote the best results.}
\caption{Comparison with state-of-the-art trackers on the 100 sequences of OTB2015. The top two results are highlighted by bold and different colors: \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.07,1.00}{blue} color.}
\caption{Success metrics (\%) of the trackers for 11 attributes. The top two results are highlighted by \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.07,1.00}{blue}.}
\caption{Precision metrics (\%) of the trackers for 11 attributes. The top two results are highlighted by \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.07,1.00}{blue}.}
\caption{Comparison results on the VOT2015 dataset. The top two results are highlighted by \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.07,1.00}{blue}.}
\caption{% Layout (a) Ball-and-stick model of the NbN lattice in perspective top view. Large green balls are niobium. Small grey balls are nitrogen. The unit cell is indicated by a solid black line. (b) Structure of the NbN $7$-layer system in the FET configuration (side view). {\color{blue}(c) Schematic picture of the planar averaged Kohn-Sham potential (V\ped{tot}, black line) for periodically repeated, charged slabs, in the absence of the exchange-correlation potential. Its components (not in scale) are highlighted with different colors: monopole potential (V\ped{mono}, red line), dipole potential (V\ped{dip}, blue line), potential barrier (V\ped{bar}, magenta line), and bare ionic potential as obtained from the periodic boundary conditions (V\ped{PBC}, green line).}}
\caption{% Integrated profiles (a) $z$-dependence of the planar average of the total electronic charge in the NbN $7$-layer system, $n_{||}$, for different values of induced charge density $\Delta n_{2D}$. The green shaded band represents the typical free carrier density values for NbN thin films \cite{ChockalingamPRB2008}. The vertical solid line indicates the position of the potential barrier; the position of the NbN slab with respect to the barrier comes from the relaxation of the supercell in the presence of the field. (b) Planar averaged difference in the full charge density between the doped and undoped slab, $\Delta n_{||}$, for different values of induced charge density $\Delta n_{2D}$. {\color{blue}Each curve is vertically offset for clarity (vertical steps are $1\cdot10^{22}\,\mathrm{cm^{-3}}$ for the three bottom curves and $5\cdot10^{22}\,\mathrm{cm^{-3}}$ for the two top curves).} Vertical dashed lines mark the positions of the atomic planes.}
\caption{% Screening (a) Number of atomic layers involved in the screening of the electric field, as a function of the induced charge density. Filled blue dots are the results of the DFT calculations. Hollow black dots are taken from the effective screening length as determined experimentally \cite{PiattiPRB2017}. {\color{blue}Filled green dots are the screening length determined with a non-linearized Thomas-Fermi approach \cite{PiattiPRB2017}, and are shown for comparison.} (b) Contributions to the screening charge due to the induced (black {\color{blue}up triangles}) and displaced (red {\color{blue}down triangles}) charge densities, and their ratio (violet {\color{blue}diamonds}), as a function of the induced charge density. {\color{blue}Solid lines are guides to the eye.}}
\caption{Page-level temporal distribution of ``manchester bombing'' \textit{General} and \textit{News} vertical SERP collections showing multiple page movement patterns (Section \ref{res:disOfStoriesPerTime}). Stories in \textit{General} SERP collections persist longer than stories in \textit{News} vertical collections. Color codes - \textcolor[RGB]{34, 185, 4}{page 1}, \textcolor[RGB]{128, 255, 104}{page 2}, \textcolor[RGB]{230, 230, 0}{page 3}, \textcolor[RGB]{109, 109, 109}{page 4}, \textcolor[RGB]{251, 0, 6}{page 5}, and blank for outside pages 1 - 5.}
\caption{Page-level temporal distribution of stories in the ``manchester bombing'' \textit{General} SERP collection showing multiple page movement patterns. Stories in \textit{General} SERP collections persist longer than stories in \textit{News} vertical collections. Color codes - \textcolor[RGB]{34, 185, 4}{page 1}, \textcolor[RGB]{128, 255, 104}{page 2}, \textcolor[RGB]{230, 230, 0}{page 3}, \textcolor[RGB]{109, 109, 109}{page 4}, \textcolor[RGB]{251, 0, 6}{page 5}, and blank for outside pages 1 - 5.}
\caption{Page-level temporal distribution of stories in the ``manchester bombing'' \textit{News} vertical SERP collection showing multiple page movement patterns, and the shorter lifespan of \textit{News} vertical URIs (compared to \textit{General} SERP URIs). Color codes - \textcolor[RGB]{34, 185, 4}{page 1}, \textcolor[RGB]{128, 255, 104}{page 2}, \textcolor[RGB]{230, 230, 0}{page 3}, \textcolor[RGB]{109, 109, 109}{page 4}, \textcolor[RGB]{251, 0, 6}{page 5}, and blank for outside pages 1 - 5.}
\caption{Temporal distributions: Stories in \text{General} SERP collections (a \& c) persist longer (``longer life'') than stories in\textit{News} vertical collections (b \& d). Compared to the ``trump russia''\textit{General} SERP collection, the stories in the ``hurricane harvey'' \textit{News} vertical collection have a ``longer life'' due to a lower rate of new stories.}
\caption{Example of a k-means clustering for the M6.5 flare on June 6, 2015. The \ion{Mg}{2} k-line profiles for a single raster are assigned to the groups in the bottom panel that they are most similar to. The quiet Sun was not colored because its profiles belong to different groups. For the full temporal evolution of this flare, see the \textcolor{blue}{online movie}.}
\caption{SJIs of four different flares in the 1400 $\text{\AA}$ channel with color coded group assignments overplotted (see Figure \ref{centroids}) and x- and y-axes given in arcseconds. Profiles assigned to group 11 (bottom right of panel C) and 12 appear at the leading edge of the flare ribbons in black. The temporal evolution of each flare can be seen in the \textcolor{blue}{online movies}.}
\caption{IRIS 1400 SJI with slanted black lines marking the positions of the raster steps. The time of the first slit position is shown in the top right corner along with the rasters cadence. The RHESSI hard X-ray contours for levels [.20, .35, .50, .65, .80, .90] appear in orange, with the noise level of the image starting at .15. An insert in the bottom left hand corner shows the 3 spectra assigned to centroid 52. The cyan and black markers can be seen to follow the leading edge of the ribbon front in the \textcolor{blue}{online movies}. For clarity, only the ribbon-front profiles are shown.}
\caption{k/h ratios for groups 0, (11+12) and 52 for all 33 flares with h\&k intensities given in absolute units. The optically thick and thin ratios are represented by the dashed lines$y=x$ and $y=2x$. Only the high intensity portion of the range $10^5-10^8$ has been plotted. Higher intensity profiles generally occur closer to the optically thick ratio, implying that during a flare, the h\&k lines remain optically thick.}
\caption{\textcolor{blue}{Approximation revenue} towards time for Airline.\label{fig:airlineConvergence}}
\caption{Illustrations of compression artifacts and soft decoding. (a) JPEG-compressed image in the case of QF = 10 (PSNR = 25.79 dB, SSIM = 0.7621, PSNR-B = 23.48 dB); (b) Soft decoded result of (a) using the developed DPW-SDNet (PSNR = {\textcolor{red}{28.22}} dB, SSIM = {\textcolor{red}{0.8376}}, PSNR-B = {\textcolor{red}{27.84}} dB).}
\caption{Average PSNR (dB)/SSIM/PSNR-B (dB) scores of different soft decoding algorithms on Classic$5$ and LIVE$1$. The best and the second-best scores are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively. }
\caption{Visual quality comparison of different soft decoding methods on \emph{Barbara} in the case of QF = 10. (a) Original image (PSNR (dB), SSIM, PSNR-B (dB)); (b) JPEG (25.79, 0.7621, 23.48); (c) CONCOLOR \cite{Zhang2016CONCOLOR} (27.73, {\textcolor{blue}{0.8216}}, 27.63); (d) D2SD \cite{Liu2016Data} ({\textcolor{blue}{27.93}}, 0.8214, {\textcolor{blue}{27.64}}); (e) ARCNN \cite{Dong2015Compression} (26.92, 0.7967, 26.75); (f) TNRD \cite{chen2017trainable} (27.24, 0.8099, 27.13); (g) DnCNN-3 \cite{Zhang2017Beyond} (27.58, 0.8161, 27.29); (h) Proposed DPW-SDNet ({\textcolor{red}{28.22}}, {\textcolor{red}{0.8376}}, {\textcolor{red}{27.84}}).}
\caption{Visual quality comparison of different soft decoding methods on \emph{Bike} in the case of QF = 10. (a) Original image (PSNR (dB), SSIM, PSNR-B (dB)); (b) JPEG (25.77, 0.7417, 23.02); (c) CONCOLOR \cite{Zhang2016CONCOLOR} (27.00, 0.7801, 27.00); (d) D2SD \cite{Liu2016Data} (27.11, 0.7859, 26.97); (e) ARCNN \cite{Dong2015Compression} (27.41, 0.7924, 27.11); (f) TNRD \cite{chen2017trainable} (27.54, 0.7971, 27.22); (g) DnCNN-3 \cite{Zhang2017Beyond} ({\textcolor{blue}{27.59}}, {\textcolor{blue}{0.7999}}, {\textcolor{blue}{27.28}}); (h) Proposed DPW-SDNet ({\textcolor{red}{28.04}}, {\textcolor{red}{0.8133}}, {\textcolor{red}{27.58}}).}
\caption{Visual quality comparison of different soft decoding methods on \emph{Lighthouse3} in the case of QF = 10. (a) Original image (PSNR (dB), SSIM, PSNR-B (dB)); (b) JPEG (28.29, 0.7636, 25.98); (c) CONCOLOR \cite{Zhang2016CONCOLOR} (29.77, 0.7976, 29.36); (d) D2SD \cite{Liu2016Data} (29.77, 0.7977, 29.24); (e) ARCNN \cite{Dong2015Compression} (29.63,0.7973, 29.19); (f) TNRD \cite{chen2017trainable} (29.75, {\textcolor{blue}{0.8013}}, 29.27); (g) DnCNN-3 \cite{Zhang2017Beyond} ({\textcolor{blue}{29.81}}, 0.8007, {\textcolor{blue}{29.38}}); (h) Proposed DPW-SDNet ({\textcolor{red}{30.30}}, {\textcolor{red}{0.8104}}, {\textcolor{red}{29.76}}).}
\caption{Average PSNR (dB)/SSIM/PSNR-B (dB) scores of different variants of the DPW-SDNet on Classic$5$ and LIVE$1$. The best scores are highlighted in {\color{red}{red}}.}
\caption{Comparisons of PSNR (dB)/SSIM/PSNR-B (dB) scores of the DnCNN-3 \cite{Zhang2017Beyond}, DPW-SDNet, and B-DPW-SDNet on Classic$5$ and LIVE$1$. The best and the second-best scores are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively.}
\caption{Our model aims to capture both: \\ \textbf{syntactic} (verb \textit{ordered} \textcolor{red}{$\rightarrow$} subj/obj \textit{boy, coffee})\\ \textbf{alignment} (noun \textit{girls} $\rightarrow$ determiner \textit{the}) attention.}
\caption{Results for translating En$\leftrightarrow$De, En$\leftrightarrow$Ru, and Ru$\rightarrow$Ar. Statistical significances are marked as $^\dagger p<0.05$ and $^\ddagger p<0.01$ when compared against the baselines and \mi/\si\when compared against the FA-NMT (no-shared). The results indicate the strength of our proposed\emph{shared-attention} for NMT.}
\caption{Pseudo-point visualization on four example videos for {\color{red}training points}, {\color{green}center bias}, {\color{blue}self-supervision}, {\color{pink}independent motion}, and {\color{cyan}person detection} (depicted as points for visualization). In general, the pseudo-points are present around the action or even follow the action. When actions are not in the frame however, as shown in the right example, pseudo-points may place automatic annotations in phantom positions.}
\caption{Time-averaged kinetic energy densities as functions of baroclinicity $\Z$ for parameter values \eqref{fixedparms} and $\Pr=0.05$, \red{$\Ek=2/300$} (top two panels), $\Pr=0.05$, \red{$\Ek=1/300$} (middle two panels), and $\Pr=0.03$, \red{$\Ek=1/500$} (bottom two panels), Full and empty symbols indicate equatorially-symmetric and -asymmetric energy components, respectively. Black circles, red squares, green triangles-up and blue triangles-down indicate the energy components $\overline E_p^{s,a}$, $\overline E_t^{s,a}$, $\widetilde E_p^{s,a}$, $\widetilde E_t^{s,a}$, respectively. Axially-symmetric and axially-asymmetric components are plotted in the left and the right panels, respectively. Vertical dash-dotted lines indicate transition points. The ranges over which distinct states are observed are indicated by arrows near the bottom abscissa, with some states co-existing as indicated. Energy components not shown are at least 10 orders of magnitude smaller than the ones shown. (Colour online) }
\caption{ Flow structures with increasing baroclinicity $\Z \times 10^{-4} = 1$, $12$, $18$, $18$ and $30$ from top to bottom for $\Pr=0.05$, \red{$\Ek=2/300$} and fixed values \eqref{fixedparms}. Bistability occurs at $\Z \times 10^{-4} = 18$. The first plot in each row shows isocontours of $\overline{u}_\varphi$ (left half) and streamlines $r\sin\theta(\upartial_\theta \overline{\pol})=$ const. (right half) in the meridional plane. The second plot shows isocontours of $u_r$ at $r=r_i+0.7$ \red{maped to the spherical surface using isotropic Aitoff projection}. The third plot shows isocontours of $u_r$ in the equatorial plane. The isocontours are equidistant with positive isocontours shown by solid lines, negative isocontours shown by broken lines and the zeroth isocontour shown by a dotted line in each plot. \red{All contour plots are snapshots at a fixed representative moment in time.} \red{Letters at each row denote corresponding flow states as indicated in Figure \ref{fig:020}.} (Colour online) }
\caption{ Flow structures with increasing baroclinicity $\Z \times 10^{-4} = 30$, $32$, $34$ for $\Pr=0.05$, \red{$\Ek=1/300$} and fixed values \eqref{fixedparms}. \red{The same quantities are plotted in each row as in Figure \ref{fig:010} and letters on the left side denote the corresponding flow state as in Figure \ref{fig:020}.} (Colour online) }
\caption{ Flow in the case $\Z \times 10^{-4} = 120$, for $\Pr=0.03$, \red{$\Ek=1/500$} and fixed values \eqref{fixedparms}. \red{The same quantities are plotted in each row as in Figure \ref{fig:010} and letters denote the corresponding flow state as in Figure \ref{fig:020}.} (Colour online) }
\caption{Summary of \red{symmetry} properties and \red{dynamo capability} of states in the case $\Pr=0.05$, \red{$\Ek=2/300$} and fixed values \eqref{fixedparms}.}
\caption{A dynamo solution \red{for a flow in state \textbf{B}} in the case $\Z \times 10^{-4} = 100$, with $\Pr=0.03$, \red{$\Ek=1/500$}, $\Pm=4$ and values \eqref{fixedparms}. First row: The first plot shows isocontours of radial magnetic field ${B}_r$ at $r=r_o+0.1$ \red{in isotropic Aitoff projection.} The second plot shows contours of $B_r$ in the equatorial plane. The third plot shows isocontours of $\overline{B}_\varphi$ (left half) and meridional field lines $r \sin\theta\, \upartial_\theta \overline{h} = $ const. (right half). Second row: The first plot shows isocontours of radial velocity $u_r$ at $r=r_i+0.7$ \red{in isotropic Aitoff projection.} The second plot shows contours of $u_r$ in the equatorial plane. The third plot shows isocontours of $\overline{u}_\varphi$ (left half) and streamlines $r\sin\theta(\upartial_\theta \overline{\pol})=$ const. (right half) in the meridional plane. \red{All contour plots are snapshots at a fixed representative moment in time.} \red{The letter B denotes the corresponding flow state as indicated in Figure \ref{fig:020}.} (Colour online)}
\caption{A dynamo solution \red{for a flow in state \textbf{B}} in the case $\Z \times 10^{-4} = 110$, $\Pr=0.03$, \red{$\Ek=1/500$}, $\Pm=2$ and values \eqref{fixedparms}. The same quantities are plotted as in Figure \ref{fig:041} \red{and the letter B denotes the corresponding flow state as indicated in Figure \ref{fig:020}.} (Colour online)}
\caption{\red{A dynamo solution \red{for a flow in state \textbf{D}} in the case $\Z \times 10^{-4} = 32$, $\Pr=0.05$, \red{$\Ek=1/300$}, and $\Pm=5$ and values \eqref{fixedparms}. The same quantities are plotted as in Figure \ref{fig:041} \red{and the letter D denotes the corresponding flow state as indicated in Figure \ref{fig:020}.} (Colour online)}}
\caption{\red{A dynamo solution \red{for a flow in state \textbf{E}} in the case $\Z \times 10^{-4} = 34$, $\Pr=0.05$, \red{$\Ek=2/300$}, and $\Pm=1.5$ and values \eqref{fixedparms}. The same quantities are plotted as in Figure \ref{fig:041} \red{and the letter E denotes the corresponding flow state as indicated in Figure \ref{fig:020}.} (Colour online)}}
\caption{\red{ (Colour online) Left panel (a): Time-averaged total magnetic energy density as a function of the magnetic Prandtl number for \red{$\Ek=2/300$} in the cases $\Pr=0.1$, $\Z=17\times10^4$ (blue triangles), $Pr=0.08$, $\Z=20\times10^4$ (black circles), $P=0.05$, $\Z=32\times10^4$ (red squares). Small black rectangles indicate the boundary of a region of no dynamo excitation for the respective sequence. Right panel (b): Components of the magnetic and kinetic energy densities (as labelled) as a function of time in the case $Pr=0.08$, $\Z=20\times10^4$, \red{$\Ek=2/300$}, $\Pm=12$. }}
\caption{\small (a). 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. (b). Live demo of the OpenNMT system.}
\caption{Overview of the distributed resource and workload model. Jobs have different resource usage patterns (computation and communication), occur at different scales (Job A and Job A'), utilize compute nodes with heterogeneous hardware specifications, and experience resource contention on a shared node (\eg, tasks \protect\includegraphics[height=6pt]{img/bluestar} and \protect\includegraphics[height=6pt]{img/redcircle}).}
\caption{A posteriori comparison between prediction of Jacobian regularized model and ground truth for POD coefficient of buoyant mixing flow. Dashed: model. Solid: ground truth. First mode: \textcolor{color1th}{$\bullet$}. Second mode: \textcolor{color2th}{$\bullet$}. Third mode: \textcolor{color3th}{$\bullet$}. Fourth mode: \textcolor{color4th}{$\bullet$}. Fifth mode: \textcolor{color5th}{$\bullet$}. Sixth mode: \textcolor{color6th}{$\bullet$}. Seventh mode: \textcolor{color7th}{$\bullet$}. Eighth mode: \textcolor{color8th}{$\bullet$}. Ninth mode: \textcolor{color9th}{$\bullet$}. Tenth mode: \textcolor{color10th}{$\bullet$}.}
\caption{We show that our part-of-speech (POS) based method achieves the trifecta of {\bf high accuracy, fast computation} and {\bf more diversity}. Beam search and diverse beam search are slow. They also produce captions with high mutual overlap and lower distinct $n$-grams than POS (see mBleu-4, div-1 and div-2 in \tabref{tab:diversity}). POS outperforms AG-CVAE on captioning metrics in \figref{fig:topk} and is therefore more accurate. {\color{blue}why is there a dash for AG-CVAE?} }
\caption{The parabolic evaluation probabilities (\ref{prob}) for {\color{red}$\gamma=1/\log{n}$} and {\color{blue}$\gamma=1/e$}.}
\caption{(a) Stability diagram of stable {\it vs.} unstable propagation front by varying the flow rate $Q$, for different gap gradients $\alpha$ (while $h_0=250$ $\mu m$). The general trend shows that for each $\alpha$, stable and complete sweep occurs at a relatively small $Q$ (denoted by $\bullet$), whereas unstable fingering propagation emerges at large $Q$ ($\circ$). For a radially-tapered cell of $\alpha= -4.75 \times 10^{-2}$, $h_0=250~\mu$m, time-evolution of top-view, stable interfaces with $Q=20$ mL/min in (b) (corresponding to \textcolor{red}{$\Box$} in (a)), while unstable interfaces in (c) for $Q=70$ mL/min (\textcolor{red}{$\Diamond$} in (a)). The time steps are $\Delta t=9$ s and $1$ s between each contours for (b) and (c), respectively.}
\caption{(a) Variation of the critical Capillary number $Ca^\ast$ separating stable {\it vs} unstable displacements for different depth gradients $\alpha$ and $h_0$. We compare the experimental values (\textcolor{blue}{$\CIRCLE$}, \textcolor{red}{$\blacksquare$}, \textcolor{green}{$\blacklozenge$}) to the theoretical $Ca^*_{th}$ (\textcolor{blue}{$\Circle$}, \textcolor{red}{$\square$}, \textcolor{green}{$\lozenge$}) derived from Eq.~(\ref{eqn2}) with $\alpha$, $r$ and $N$ from our experimental results and parameters. (b) Surface plot of theoretical $Ca^*_{th}$ greatly depends on $r$ and $N$ using Eq.~(\ref{eqn2}), for $\alpha= -8.66 \times 10^{-2}$ and $h_0=250$ $\mu m$, showing a stable displacement when $Ca < Ca^\ast$ whereas unstable one when $Ca > Ca^\ast$.}
\caption{Green curves show the first unity beta surfaces (solid line for $\beta_p=1$; dashed line for $\beta_{p+e}=1$) computed from the model (Run II-C) superimposed on a \textit{STEREO} image from \cite{deforest2016ApJ828}. White `+' shows location of enhanced turbulence inferred by \citet{lotova1985AA150} (see Figure \ref{fig:scintillation}); \textit{Helios} perihelion is shown as `{\color{blue}$\oplus$}'; the first three perihelia of the \psp are shown as `$\otimes$'.}
\caption{Enhanced scintillation ($m\nu$) region from the observations of \cite{lotova1985AA150}, seen as a bump at \(\sim 20~\rs\) in the dashed red curve. Radial solar wind speed \(V_r\) (dash-dotted blue curve) and turbulence energy density (per unit mass) \(Z^2\) (solid black curve) are shown at (a) an ecliptic heliolatitude of 6.75\degree~and (b) a polar heliolatitude of 82\degree. Panel (a) shows shaded bands representing the locations of the Alfv\'en (pale blue band) and sonic (grey band with dashed outline) surfaces in the ecliptic region of the simulation (between heliolatitudes 6.75\degree~and \(- 6.75\degree\)). Panel (b) shows vertical lines representing locations of the Alfv\'en surface (pale blue solid) and the sonic surface (grey dashed) at 82\degree~heliolatitude. All simulation results shown here are from Run I-A. The first, third, and final perihelia of the \psp are represented as $\oplus$ symbols, {at heliocentric distances of 35.66, 20.35, and 9.86\(~\rs\), respectively \citep{fox2016SSR}.}}
\caption{\psp crosssings of the critical surfaces are illustrated by a sequence of meridional planes that contain the spacecraft trajectory. The 8th orbit is depicted in a 10\degree~dipole simulation (Run I-C; see Figure \ref{fig:barplot2}(a)), representing solar-minimum conditions. The sonic, Alfv\'en, and first (proton+electron) beta unity surfaces are depicted as solid pink, solid blue, and dashed green curves, which are superposed on contours of proton density. The\psp position is at the center of the yellow `+' symbol. A video animation is available as Supplementary Material.}
\caption{Qualitative comparison of our approach trained on generated and real data with state-of-the-art trackers, \textcolor{cyan}{CREST}, \textcolor{magenta}{TCNN}, \textcolor{blue}{EBT} and \textcolor{green}{DSLT} on the some challenging sequences, \textit{excavator}, \textit{jacket}, \textit{mixed\_distractors}, {\textit{garden}}, {\textit{quadrocopter2}}, {\textit{boat2}}, {\textit{bird}} and \textit{trees2} in VOT-TIR2017~\cite{Kristan_2017_ICCV}. Yellow dashed bounding box means \textcolor{yellow}{Groundtruth} and red solid bounding box is \textcolor{red}{Ours}. {The last two rows show failure cases of our tracker}. }
\caption{Additional \aastex\symbols}
\caption{\red{A demonstration of the progression of cell division from one cell into sixteen cells at a constant total cell volume. The underlying spatial mesh/discretization over the simulation domain is shown in (a). Cell interiors are shown in red and the cell membrane in cyan-yellow.}}
\caption{Evolution of total free energy with time (normalized). \red{Each spike in the energy curve corresponds to transient repulsion between newly formed daughter cells following a cell division. The corresponding cell division events are shown in the inset sub-figures.}}
\caption{Evolution of the total free energy with time (normalized). \red{Each spike in the energy curve corresponds to transient repulsion between newly formed daughter cells following a cell division, and the corresponding cell division events are shown in the inset sub-figures.}}
\caption{\red{Comparison of cell morphologies during early embryogenesis observed in (a) nematodes (\emph{C. elegans} and \emph{Prionchulus sp.}), and (b) the computational model.}}
\caption{\red{Early embryogenesis of nematodes up to the eight-cell stage observed in \emph{C. elegans}, (a-c), and \emph{Prionchulus sp.}, (g-i). Some of the cells are either partially or completely hidden from view in (c) and (i). Figure reproduced from Schierenberg \cite{Schierenberg2006} (Original figure distributed under Creative Commons Attribution License).}}
\caption{\red{Computations showing up to the third generation of divisions resulting in eight cells at a constant total cell volume.}}
\caption{\textbf{Scene flow on KITTI point clouds.} We show scene flow predicted by FlowNet3D on four KITTI scans. Lidar points are colored to indicate points as from \textcolor{red}{frame 1}, \textcolor{green}{frame 2} or as \textcolor{blue}{translated points} (point cloud 1 + scene flow).}
\caption{Visualization of local point feature similarity. Given a \textcolor{magenta}{point $P$} (pointed by the blue arrow) in frame 1 (gray), we compute a heat map indicating how points in frame 2 are similar to $P$ in feature space. More red is more similar.}
\caption{The error rates(\%) of CIFAR-10 and CIFAR-100. Best results are written in {\color{blue}{\textbf{blue}}}. The character * indicates results are directly obtained from the original paper.}
\caption{The error rates (\%, 1-crop testing) on ImageNet. Best results are written in {\color{blue}{\textbf{blue}}}. The character * indicates results are directly obtained from the original paper.}
\caption{The error rates (\%) on PAMAP2 and Fashion-MNIST. Best results are written in {\color{blue}{\textbf{blue}}}.}
\caption{ \textit{Left}: Inference in \gls{AIR}. The \textcolor{pink}{pink \gls{RNN}} attends to the image sequentially and produces one latent variable $\bzt^i$ at a time. Here, it decides that two latent variables are enough to explain the image and $\bzt^3$ is not generated. % \textit{Right}: Inference in \gls{SQAIR} starts with the \gls{prop} phase. \gls{prop} iterates over latent variables from the previous time-step $t-1$ and updates them based on the new observation $\bxt$. The \textcolor{blue}{blue \gls{RNN}} runs forward in time to update the hidden state of each object, to model its change in appearance and location throughout time. The \textcolor{orange}{orange \gls{RNN}} runs across all current objects and models the relations between different objects. Here, when attending to $\bz^1_{t-1}$, it decides that the corresponding object has disappeared from the frame and \textit{forgets} it. Next, the \gls{disc} phase detects new objects as in \gls{AIR}, but in \gls{SQAIR} it is also conditioned on the results of \gls{prop}, to prevent rediscovering objects. See \Cref{fig:sqair_inf_detail} for details of the colored \glspl{RNN}.}
\caption{ \textit{Left}: Interaction between \gls{prop} and \gls{disc} in \gls{SQAIR}. Firstly, objects are propagated to time $t$, and object $i=7$ is dropped. Secondly, \gls{disc} tries to discover new objects. Here, it manages to find two objects: $i=9$ and $i=10$. The process recurs for all remaining time-steps. \textcolor{blue}{Blue arrows} update the temporal hidden state, \textcolor{orange}{orange ones} infer relations between objects, \textcolor{pink}{pink ones} correspond to discovery. \textit{Bottom}: Information flow in a single discovery block (\textit{left}) and propagation block (\textit{right}). In \gls{disc} we first predict \textit{where} and extract a glimpse. We then predict \textit{what} and \textit{presence}. \Gls{prop} starts with extracting a glimpse at a candidate location and updating \textit{where}. Then it follows a procedure similar to \gls{disc}, but takes the respective latent variables from the previous time-step into account. It is approximately two times more computationally expensive than \gls{disc}. For details, see \Cref{algo:sqair_prop,algo:sqair_disc} in \Cref{app:algo}. }
\caption{The average overlap score of nine compared methods for 15 sequences. Bold number in \textcolor{blue}{blue} indicates the best performance, while \textcolor{red}{red} indicates the second best.}
\caption{Tracking results of different methods. Frame indexes are shown at the top left corner of each figure.\\ (\textcolor{black}{\textbf{---}}L1T, \textcolor{green}{\textbf{---}}MTT, \textcolor{StruckCol}{\textbf{---}}Struck, \textcolor{MILCol}{\textbf{---}}MIL, \textcolor{yellow}{\textbf{---}}IVT, \textcolor{cyan}{\textbf{---}}VTD, \textcolor{Brownish}{\textbf{---}}MTMVTLS, \textcolor{red}{\textbf{---}}MTMVTLAD, \textcolor{blue}{\textbf{---}}SMTMVT)}
\caption{ {\bf Network Architecture.} The \boldyellow{feature encoder} takes a raw LIDAR point cloud as input, groups the points into $W\times H\times 10$ voxels, and outputs 128D feature maps (for clarity, the size of the feature maps is not shown in the figure) which are concatenated and passed to the \boldgreen{context encoder}. The context encoder learns a global representation by interleaving convolution with strided convolution layers and ``flattening'' the third dimension (height above ground), \ie, we assume that 3D objects cannot be located on top of each other and that 3D scene points that project to the same location in the ground plane undergo the same 3D motion. Feature maps at different resolutions are upsampled, stacked and fed into the decoding branches. 3D scene flow is computed for every input voxel in the \boldorange{scene flow decoder} and the result is passed to the \boldred{rigid motion decoder}, which infers a rigid body transformation for every point. In parallel, the \boldcyan{ego-motion regressor}, further downsamples the feature map by interleaving convolutional layers with strided convolutional layers and a fully connected layer at the end to regress rigid motion for the ego vehicle. In addition, the \boldpurple{object decoder} predicts the location and size (\ie, 3D bounding box) of objects in the scene. Finally, the \boldblue{object motion decoder} takes the point-wise rigid body motions as input and predicts the object rigid motions by pooling the rigid motion field over the detected 3D objects. }
\caption{{\bf Rigid Motion Estimation.} In (\subref{fig:equivariance_illustration}), indices $A$ and $B$ denote the coordinate system of points $\bp$ and $\bq$ at origin $\vec{o}_A$ and $\vec{o}_B$, respectively. The same scene flow $\bv$ can locally be explained with the same rigid body motion $(\bR_L,\bt_L)$, but requires different translations $\bt^\vec{p}_{W}\neq\bt^\vec{q}_{W}$ in the global coordinate system. A simple example (\subref{fig:equivariance_results}) provides empirical evidence that translation cannot be learned in global coordinates with a CNN. Using global coordinates, the translation error increases significantly with the magnitude of rotation (\textcolor{darkgreen}{green}). There is no such increase in error when using local coordinates (\textcolor[rgb]{1,0.5,0.3}{orange}).}
\caption{(a) Device schematic and measurement configuration, including load resistance, DC and noise measurement circuits connected to contact 1. (b) Shot noise measurements versus bias current. The symbols represent shot noise, measured in configuration (a) at temperatures (top to bottom) 800\,mK (crosses), 600\,mK (triangles), 400\,mK (squares), 200\,mK (hexagons) and 80\,mK (diamonds). The dashed line is the shot noise prediction with$F=0.29$ at 80\,mK. (c) Two-terminal linear-response resistance of a 5\,$\rm\mu m$ long edge 1-2 as a function of temperature obtained in a different cooldown.}
\caption{Quantitative evaluation of trackers under different tracking challenges using AUC(\%) of success plot on OTB-50. The {\color{red}first}, {\color{green}second} and {\color{blue}third} best results are shown in color.}
\caption{Image and textual attention visualization on val unseen (best viewed at 200\%). At each step, the textual attention is shown at the top, and the 1st, 2nd and 3rd most attended view angles are shown in \textcolor{red}{red}, \textcolor{orange}{orange} and \textcolor{yellow}{yellow} boxes, respectively (the number in the parenthesis shows the attention weight). The red arrow shows the direction chosen by the agent to go next.}
\caption{Image and textual attention visualization on val unseen (best viewed at 200\%). At each step, the textual attention is shown at the top, and the 1st, 2nd and 3rd most attended view angles are shown in \textcolor{red}{red}, \textcolor{orange}{orange} and \textcolor{yellow}{yellow} boxes, respectively (the number in the parenthesis shows the attention weight). The red arrow shows the direction chosen by the agent to go next.}
\caption{Image and textual attention visualization on val unseen (best viewed at 200\%). At each step, the textual attention is shown at the top, and the 1st, 2nd and 3rd most attended view angles are shown in \textcolor{red}{red}, \textcolor{orange}{orange} and \textcolor{yellow}{yellow} boxes, respectively (the number in the parenthesis shows the attention weight). The red arrow shows the direction chosen by the agent to go next.}
\caption{Terminology defined for different types of deposits observed in this study.}{\includegraphics[width=\textwidth]{terminology}}
\caption{Stratigraphic features of selected tephras in Sequences E-C at TS, northwest of Mono Lake. }{\includegraphics[width=\textwidth]{TS_e2c}}
\caption{Stratigraphic features of selected tephras in Sequences D-C at SEB, southeast of Mono Lake.}{\includegraphics[width=1.2\textwidth]{SEB_d2c}}
\caption{Stratigraphic features of selected tephras in Sequence B at SEB, TS, WS, and CC.}{\includegraphics[width=1.2\textwidth]{SEB_b}}
\caption{Stratigraphic features of A2 at HMC, RC1, RC2, ND1, and WS. Yellow cells represent the correlated sub-unit from the most voluminous eruption pulse of the A2 eruption.}{\includegraphics[width=1\textwidth]{a2_summary}}
\caption{Established sub-unit correlation of tephras in Sequences C-B between sample sites with evidence for correlation.}{\includegraphics[width=1.2\textwidth]{cor_b}}
\caption{Established sub-unit correlation of tephras in Sequence A between sample sites with evidence for correlation. Correlations are made separately for sample sites in the east and west of the Mono Craters.}{\includegraphics[width=1.2\textwidth]{cor_a_linux}}
\caption{Effect of neutrally buoyant particles on the laminar-turbulent transition threshold, as depicted from the temporal evolution of $C_f$, after decreasing $\Rey_b$ in $(a)$ Couette flow and $(b)$ pressure-driven flow. The initial flow configuration of the Couette (resp. channel) flow is taken from a fully-turbulent simulation at $\Rey_b=500$ (resp. 2300). $(a)$: \figline $\Rey_b=500$, $\mathit{\Phi}=5\%$; \textcolor{green}{\figldash} $\Rey_b=470$, $\mathit{\Phi}=10\%$; \textcolor{red}{\figddash} $\Rey_b=455$, $\mathit{\Phi}=10\%$; \textcolor{cyan}{\figsdash} $\Rey_b=440$, $\mathit{\Phi}=10\%$; \textcolor{blue}{\figddash} $\Rey_b=455$ to $345$, $\mathit{\Phi}=5\%$ and $I$ to $V$ corresponding to $\Rey_b=455$, $415$, $390$, $365$ and $355$. $(b)$: \figline $\Rey_b=2000$, $\mathit{\Phi}=1\%$; \textcolor{blue}{\figldash} $\Rey_b=2000$, $\mathit{\Phi}=5\%$; \textcolor{red}{\figddash} $\Rey_b=1700$, $\mathit{\Phi}=5\%$; \textcolor{green}{\figsdash} $\Rey_b=1500$, $\mathit{\Phi}=5\%$.}
\caption{Particle effect on flow stability. \underline{\textit{Left panel}}: Couette flow starting from a fully turbulent regime. The turbulent state is stable under single-phase condition, even when the streamwise velocity perturbations are suppressed, the flow recovers its fully turbulent nature. Adding particles damp the velocity fluctuations and make the flow laminar. \figline $\Rey_b=430$, single-phase flow removes $u'$; \textcolor{blue}{\figldash} $\Rey_b=430$, $\mathit{\Phi}=5\%$ and $L_y/d=10$; \textcolor{red}{\figddash} $\Rey_b=455$, single-phase flow removes $u'$; \textcolor{green}{\figsdash} $\Rey_b=455$, $\mathit{\Phi}=10\%$ and $L_y/d=20$. \underline{\textit{Right panel}}: channel flow starting with a flow distribution according to (\ref{eq:artificial_1}) where an artifical streak is initially imposed. The single-phase flow tends towards the laminar state at $\Rey_b=2600$ which is above the laminar-turbulent transition. Adding small number of particles in the flow triggers the transition to turbulence. \figline single-phase flow; \textcolor{blue}{\figldash} $\mathit{\Phi}=0.5\%$ and $L_y/d=16$; \textcolor{red}{\figddash} $\mathit{\Phi}=0.75\%$ and $L_y/d=16$; \textcolor{green}{\figsdash} $\mathit{\Phi}=0.5\%$ and $L_y/d=20$. }
\caption{Particle distribution $x-z$ plane. $(a)$ and $(b)$ are the concentration contours taken at $y/L_y=0.2$ in $x-z$ plane averaged over $80$ time units for $C500-5$ and $P2600-5$, the isolines show $u'/U_{bulk}$ in $x-z$ plane where dashed line stands for negative and solid line shows positive, the interval is $0.04$ in $(d)$ and $0.03$ in $(e)$. The profiles of $u'/U_{bulk}$ averaged in streamwise are also plotted on the right side of $(a)$ and $(b)$, separately. $(c)$ and $(d)$ are the mean velocity of \figline fluid phase, \textcolor{blue}{\figldash} particle phase and \textcolor{red}{\figddash} difference between fluid phase with particle phase.}
\caption{ Top panels: Couette flow; Bottom panels: channel flow. $(a)$ and $(d)$ show the one-dimensional streamwise and spanwise wavenumber energy spectra of the streamwise velocity $E_{uu}$ averaged in the wall-normal direction. In $(a)$ \figline $C500-0$; \textcolor{blue}{\figldash} $C500-5$; \textcolor{red}{\figddash} $C500-10$. In $(d)$ \figline $P2600-0$; \textcolor{blue}{\figldash} $P2600-5$; \textcolor{red}{\figddash} $P2600-10$. Contour figures show the two-dimensional contours of the energy spectra. $(b)$ and $(e)$: single-phase flow with $C500-0$ and $P2600-0$. $(c)$ and $(f)$: two-phase flow with $C500-5$ and $P2600-5$.}
\caption{Simultaneous temporal evolution of the friction coefficient $C_f$ and the near wall streamwise vorticity, in Couette flow ($(a)$ and $(b)$), and channel flow ($(c)$ and $(d)$). $(a)$ and $(c)$ plot the summation of the amplitude square of x-independent vortices ($m=0$) in different spanwise wavenumbers ($1\leq n \leq N_z/2$) and integrated in the near wall region ($y^+<15$). $(b)$ and $(d)$ show x-independent vortices ($m=0$) in the near wall region ($y^+<15$) as a function of spanwise wavelength ($2L_z/N_z \leq \lambda_z \leq L_z$). The line style indicates single-phase (solid) and two-phase (dashed line) flows. In $(a)$ and $(b)$: \textcolor{blue}{\figline} $C500-0$; \textcolor{red}{\figldash} $C500-5$, in $(c)$ and $(d)$: \textcolor{blue}{\figline} $P2600-0$; \textcolor{red}{\figldash} $P2600-5$.}
\caption{$(a)$ and $(b)$ show temporal evolution of \figline circulation (integrated over $0.1<y/L_y<0.4$) and \textcolor{red}{\figsdash} spacial average of absolute value of vorticity stretching ($\mid \omega_x\partial u/\partial x \mid$ within $0.1<y/L_y<0.4$ where large scale vortices take place as seen in figure \ref{fig:Figure_5_1}(b)). $(c-f)$ show contours of the streamwise velocity fluctuations $u'/U_{bulk}$ in the snapshot plane at $y/L_y=0.2$, showing the streaks. The interval of isolines is 0.04. \figsdash stands for $u'/U_{bulk}<0$ and \figldash stands for $u'/U_{bulk}>0$. The color contours indicate the stretching term $\omega_x\partial u/\partial x$. Left panels are for single-phase flow $P2600-0$, two points $(c)$ and $(e)$ in $(a)$ are x-independent flow (at the trough of the vorticity stretching instant $\Delta tU_{bulk}/h=900$) and for x-dependent flow (at the peak of the vorticity stretching instant $\Delta tU_{bulk}/h=970$). Right panels are for suspension flow $P2600-1$, two points $(d)$ and $(f)$ in $(b)$ are corresponding to $\Delta tU_{bulk}/h=405$ and $\Delta tU_{bulk}/h=480$. Line styles in $(d)$ and $(f)$ are same as in $(c)$ and $(e)$.}
\caption{Benchmark image denoising results. Training and testing protocols are followed as in~\cite{tai2017memnet}. Average PSNR/SSIM for various noise levels on 14 images, BSD200 and Urban100. \textcolor{red}{Red} is the best and \textcolor{blue}{blue} is the second best performance. }
\caption{\textbf{First row:} The color code is {\color{py-3-1} blue: iplb}, {\color{py-3-2} orange: projected}, {\color{py-3-3} green: Lagrange} with \emph{continuous} lines for no approximation and \emph{dashed} lines for piecewise linear approximation. The gray dotted line is the baseline and the dashed black line marks unconstrained logistic regression. \textbf{Second row:} \emph{Continuous/dotted} lines correspond to $z=0$ and \emph{dashed/dash-dotted} lines to $z=1$. The color code is ({\color{py-4-1}red: no approx. + float}, {\color{py-4-2}purple: no approx. + fixed}, {\color{py-4-3}yellow: pw linear + float}, {\color{py-4-4}turquoise: pw linear + fixed}, gray: baseline).}
\caption{We plot the fraction of people with $z=0$ (\emph{continuous/dotted}) and with $z=1$ (\emph{dashed/dash-dotted}) who get assigned positive outcomes over the constraint $c$ for 5 different datasets. The different colors correspond to ({\color{py-4-1}red: no approximation + floats}, {\color{py-4-2}purple: no approximation + fixed-point}, {\color{py-4-3}yellow: piecewise linear + floats}, {\color{py-4-4}turquoise: piecewise linear + fixed-point}, gray: baseline).}
\caption{Test set accuracy over the $p\%$ value for different optimization methods ({\color{py-3-1} blue: iplb}, {\color{py-3-2} orange: projected}, {\color{py-3-3} green: Lagrange}) and either no approximation (\emph{continuous}) or a piecewise linear approximation (\emph{dashed}) of the sigmoid using floating-point numbers. The gray dotted line is the baseline (see Section~\ref{subsec:current}) and the black dashed line is unconstrained logistic regression (from scikit-learn).}
\caption{The fraction of people with $z=0$ (\emph{continuous/dotted}) and $z=1$ (\emph{dashed/dash-dotted}) who get assigned positive outcomes ({\color{py-4-1}red: no approx. + float}, {\color{py-4-2}purple: no approx. + fixed}, {\color{py-4-3}yellow: pw linear + float}, {\color{py-4-4}turquoise: pw linear + fixed}, {gray: baseline}).}
\caption{\footnotesize Results for all configurations and the typological profile of the 21 Europarl languages. All languages are Indo-European, except for those marked with $^*$ which are Uralic. Morpholical counting complexity (MCC) is given for each language, along with bits per English character (BPEC) and the $\Delta$BPC, which is BPEC minus bits per character (BPC). This is \textcolor{darkblue}{blue} if BPEC $>$ BPC and \textcolor{darkred}{red} if BPEC $<$ BPC.}
\caption{Two unanswerable questions written by crowdworkers, along with plausible (but incorrect) answers. Relevant keywords are shown in \textcolor{blue}{blue}. }
\caption{Weighted sigmoid cross-entropy (lower is better) for human keypoint localisation on miniKinetics test set for zero \predictionlag. ``Cl" denotes models with multi-rate clocks, ``T" -- models with temporal filters, ``FB" -- models with feedback. \textbf{Left}: Comparison between Par-Inception and Par-DenseNet for different levels of parallelism. Note that in terms of number of sequential convolutions, 14 subnetworks for Par-DenseNet are equivalent to 10 subnetworks for Par-Inception, and similar for 7(5). \textbf{Right}: Variations of Par-DenseNet. In the absence of parallelisation (1 subnetwork), the accuracy of the best models with multi-rate clocks is just slightly worse to that of a much slower sequential model. Parallelisation penalises the accuracy of models with clocks more. The basic Par-DenseNet can have up to 4 parallel subnetworks with modest drop of accuracy.}
\caption{Comparison between the weighted sigmoid cross-entropy (lower is better) of models with different levels of parallelism and the same models distilled from sequential for human keypoint localisation on miniKinetics test set for zero \predictionlag. Results presented for a DenseNet model with multi-rate clocks (``Cl"), temporal filters (``T"), and feedback (``FB"). See text for details.}
\caption{Absolute WLSE for TPB+PMMA~\cite{Benson}, PEN+PMMA, standalone PEN and PEN+glass, and relative efficiencies w.r.t. TPB+substrate. Values indicated with asterisks are directly measured in this work; the rest relies on literature based inference (see text for details). The last row shows experimental ratios of fluorescence yield of VM2000 and approx.~1~mg/cm\squared\thick TPB coatings evaporated on Tetratex\textsuperscript{\textregistered} (TTX), with the 87~K value measured with an LAr detector from Ref.~\cite{Baudis}. }
\caption{\emph{In-situ} intercalation of arc discharge SWCNTs. Steps marked with {\color{black}red} color indicate measurement steps, while {\color{OliveGreen}green} parts show intercalation steps proceeded at $545$ K. In the {\color{black}first step} the material clearly shows a semiconducting behavior (resistivity decreases upon increasing temperature) as expected from a bundled branch of nanotubes, where $2/3$ of the tubes are non-metallic. In the {\color{OliveGreen}second step} the first intercalation takes place, which lasted for $16$ minutes. This is followed by the measurement denoted with {\color{black}III}, where the system was cooled from $545$ K down to $155$ K and back. On this part the SWCNTs show a completely metallic behavior (resistivity decreases upon decreasing temperature and increases upon increasing) proving that the intercalation took place. Further doping steps, {\color{OliveGreen}IV and VI} and measurement parts, {\color{black}V and VII} indicate the drop of the resistivity as the material is turning more and more metallic.}
\caption{High Level of CBS-DL ({\color{red}and MA-DBS})}
\caption{Comparison of the computational efficiency of the Monte Carlo and dynamical methods for evaluating the noise performance of the Haus modelocking equation. We integrate the system for 15000 round trips on each simulation run of the Monte Carlo method. The tests are coded in Matlab{\textregistered} which have a memory overhead of 500\,MB that is included in the memory usage.\label{tab:costs-hmn}}
\caption{\textbf{Visualization of two-way one-shot classification} trained on synthesized examples. \textit{Correctly} classified images are framed in \textcolor{magenta}{magenta (Golden retriever)} and \textcolor{yellow}{yellow (African wild dog)}. The only two images seen at training time and used for sample synthesis are framed in \textcolor{blue}{blue}. Note the non-trivial relative arrangement of examples belonging to different classes handled successfully by our approach. The figure is plotted using t-SNE applied to VGG features. Best viewed in color.}
\caption{(Color online) Nonequilibrium steady state quantum coherence $|\rho^{ss}_{12}|$ within the two{\color{red}-}reservoir setup ($\gamma_M=0$) (a) by tuning noise-induced transition coefficients $\gamma^L_{12}$ and $\gamma^R_{12}$ with $T_L=2$ and $T_R=1$, and (b) by tuning the left and right temperatures $T_L$ and $T_R$ with ${\Delta}T=T_L-T_R$, $\gamma^L_{12}=\sqrt{\gamma^L_{11}\gamma^L_{22}}$ and $\gamma^R_{12}=0$. The other system parameters are given by $\gamma^{L(R)}_{11}=\gamma^{L(R)}_{22}=0.01$ and $\varepsilon_L=\varepsilon_R=1$.}
\caption{ Illustration of a graph matching problem matching nose and left/right feet of two penguins. The \textcolor{blue}{blue nodes} on the left penguin correspond to the underlying node set $V$, while the \textcolor{blue}{blue nodes} on the right penguin correspond to the labels $\SL$. The \textcolor{green}{green lines} denotes the matching. Note that no two labels are matched twice. The \textcolor{red}{red springs} denote pairwise costs $\theta_{ij}$ that encourage geometric rigidity of the matching. }
\caption{Crystal-field theoretical ordering of the orbitals and the orbital occupation characteristics of the ground state and first ({\color{red}{red arrow}}) and third ({\color{blue}{blue arrow}}) lowest $d-d$ electronic transitions.}
\caption{$R(t)$ diagrams for shells in the experimental device filled with deuterium which produced the record high compression. \newline (a) x-Ray images of the shell 2 (see Fig.~\ref{fig:S2}) as a function of time, where $t_0$ is the initial state, $t_1$ and $t_2$ are the compression phases, $t_3$ is the moment of maximum compression (``bounce'') and $t_4$ is the expansion phase; light and red circles are the outer and inner boundaries of the shell 2, respectively. (b) experimental data and calculated $R(t)$ diagrams, symbol \Square\denotes the results of electrocontact technique, symbols$+$ and $\times$ are used for x-ray data from the model experiment, and symbols $\Diamond$ and \Circle -- for the data of the basic experiment. }
\caption{The background extraction regions, A and B, overlaid on the Galactic diffuse \gray{} model count map centered around 9 GeV. The bright \gray{} emitting structure between these two regions is the R CrA MC.}
\caption{The background extraction regions 2, 3, and 4, overlaid on the Galactic diffuse \gray{} intensity at 9 GeV. The bright \gray{} emitting structure in the region 1 is the $\rho$ Oph MC. Grid lines of latitude parallels and longitude meridians at $30^{\circ}$ intervals are shown.}
\caption{The SED of the R CrA MC derived for the background model B and contaminated by \gray{} emission from the southern Fermi bubble is shown along with the spectral model describing the stacked spectrum of the Gould Belt MCs. The latter model is scaled down by a factor of 53.}
\caption{\redmark{ServeNet architecture}}
\caption{Accuracy values of competing models when the training data used is sparse. \textbf{Bold-faced} values are the best accuracies in the column, while \textcolor{red}{red} values are accuracies worse than NSC(LA).}
\caption{Classification accuracies of competing models. \textbf{C} refers to the additional context, $N$ refers to the number of translations. In TopCNN, word refers to using word-specific topic while sentence refers to using sentence-specific topic. %The first seven models are models without additional context. The rest of the models are with additional context. Accuracies colored \textcolor[rgb]{1,0,0}{red} are accuracies that perform worse than CNN. Previous state of the art results and the results of our best model are \textbf{bold-faced}. The winning result is \underline{underlined}. The number inside the parenthesis indicates the increase from the base model, CNN.}
\caption{Accuracies of the worst CNN+translation classifiers when $N=1$. Accuracies less than CNN accuracies are highlighted in \textcolor[rgb]{1,0,0}{red}.}
\caption{Two examples of self usability of Korean sentences from the MR data set. Texts colored in \textcolor[rgb]{1,0,0}{red} are mistranslated texts.}
\caption{\label{tab:devs} Constitutive parameters of 5 devices measured in the experiment: Josephson inductance of largest junction ($L_J$), number of SNAILs ($M$), junction \blue{inductance} ratio ($\alpha$), coupling capacitance to the $50 \,\Omega$ transmission line ($C_c$), and frequency of the $\lambda/2$ microstrip embedding structure when the array of SNAILs is replaced by a short ($\omega_0$).}
\caption{% (a) Premultiplied vorticity spectra of the simulations at the two times used as references. The wavelength, $\ell=2\pi/k$, is defined from the wavenumber magnitude $k$. \solid, $t=0$; \dashed, $\omega'_0 t= 11$; \chndot, $\omega'_0 t=26$. The line with symbols is the premultiplied energy spectrum at $\omega'_0 t= 11$. % (b) Velocity and vorticity magnitude of the reference field at $\omega'_0 t=11$. The overlaid grid is used to select the test subsets. % (c) As in (b), for $\omega'_0 t=26$. % }
\caption{% (a) P.d.f.s of the ratio of mean vorticity to mean rate of strain over individual significant cells, for 192 initial conditions with $n=1,\, N=36,\, \omega'_{11}(t-t_{11}) =3.4$. The vertical dashed line is the global average that separates vortex- from strain-dominated regions, $\bra\omega^2\ket/\bra S^2\ket=2$. \solid, Ten most significant cells for each initial condition; \dashed, ten least significant cells; \chndot, random cells. Lines without symbols are initialised using \r{eq:delome1}. Those with symbols use \r{eq:delu}. % (b) As in (a), for the kinetic energy, $q^2=u^2+v^2$. % }
\caption{(a) Number of tests required for: \dashed, full search; \circle, tree search; \solid, estimate in \r{eq:tests}. (b) Comparison of the results of the tree search (symbols), and the full search (crosses). \circle, $n=3$; \trian, $n=7$. Solid lines are the highest deviation limit, and dashed ones are the lowest. In all cases, $N=36$, the pruning size is $M=15$, and the branching number is $b=3$. }
\caption{\textcolor{red}{#1}}
\caption{Left: Tracer-density profiles (dashed black for clusters, dotted red for galaxies) around cluster-defined voids of radius $190h^{-1}$Mpc $<\rv<220h^{-1}$Mpc in the \magneticum\simulation. Solid black and long-dashed red lines show the best fits obtained via Equation~(\ref{eq:prof}). Right: Cluster- and galaxy-density profiles from the left panel plotted against each other (black points with error bars). The dotted black line shows the best fit using Equation~(\ref{eq:lin-fit}).}
\caption{Comparison of the best-fit $\bs$ obtained from our largest void sample (solid red line) to the relative bias $\lrb$ between clusters and galaxies in the \magneticum\simulation, calculated using the estimators as indicated (black dashed and dotted lines). The stellar-mass cut for the galaxy sample is varied from left to right, with the same values as in Fig.~\ref{fig:magn-bias-vs-size}.}
\caption{The abundance of voids identified in the galaxy and cluster samples of the \mice\mocks, as a function of their effective radius. Both photometric and spectroscopic redshifts have been used in each case, as indicated in the figure legend. The cluster-void size function is not significantly affected by photo-z uncertainty. In fact, clusters provide the most accurate photometric redshift measurements and cluster-voids are the largest voids, further reducing the relative impact of photo-z scatter on void finding.}
\caption{Tracer-density profiles (solid black for \redmapper\clusters, dashed red for\redmagic\galaxies) around cluster-defined voids of size$50h^{-1}$Mpc $<\rv<60h^{-1}$Mpc in the \mice\mocks. The luminosity cut for the galaxy sample is varied from left to right, as indicated in each panel.}
\caption{Abundance of voids as a function of their effective radius, identified in the distribution of \redmapper\clusters from DES data (Y1A1). The average cluster-density profile of all voids is shown as inset.}
\caption{Density plot of \redmapper\clusters and their associated void centres (cyan circles) in a redshift slice of$0.2<z<0.45$. The blue line displays the 5-year-DES footprint, voids intersecting with the survey mask are discarded. }
\caption{Three-dimensional map of the DES light cone; magenta dots show $5\%$ of all \redmapper\clusters, green dots display$5\%$ of \redmapper\clusters inside watershed voids and black spheres of radius$\rv$ represent the spherical volume of each void.}
\caption{Tracer-density profiles (solid black for \redmapper\clusters, dashed red for\redmagic\galaxies) around cluster-defined voids of size$40h^{-1}$Mpc $<\rv<80h^{-1}$Mpc in the DES data. The luminosity cut for the galaxy sample is varied from left to right, as indicated in each panel.}
\caption{Comparison of Alice \ion{O}{1}~1356~\AA\emission from rows 13--17 (blue stars) with the warm (5--100~eV) electron density (red triangles) as defined by\citet{JGRA:JGRA52830} and \citet{broiles2016statistical} from the IES instrument. The x axis begins at the start of the CME as reported by the RPC magnetometer in \cite{doi:10.1093/mnras/stw2112}. \ion{O}{1} 1356 \AA\measurements that may indicate a possible outburst are marked.\textcolor{black}{This possible outburst time coincides with slightly elevated electron fluxes that may be indicative of an outburst as well}. Labels \textbf{1} and \textbf{2} denote the two electron spikes that coincide with spikes in \ion{O}{1}~1356~\AA\emission. Spectra with an off nadir angle less than 1$^{\circ}$ are used to create the plot.}
\caption{Examples of error types observed in the qualitative analysis - \textcolor{blue}{blue} indicates ground truth}
\caption{The pipeline of our unified RNN-CNN method to predict \red{and interpret} compound-protein affinity.}
\caption{Comparing the novel representations to the baseline based on RMSE (and Pearson correlation \red{coefficient r}) of pIC$_{50}$ shallow regression.}
\caption{Under novel representations learned from seq2seq, comparing random forest and variants of separate RNN-CNN and unified RNN-CNN models based on RMSE (and Pearson correlation \red{coefficient $r$}) for pIC$_{50}$ prediction.}
\caption{\red{Under novel representations learned from seq2seq, comparing different attention mechanisms of unified RNN-CNN models based on RMSE (and Pearson correlation \red{coefficient $r$}) for pIC$_{50}$ prediction.}}
\caption{\red{Comparing strategies to generalize predictions for four sets of new protein classes: original random forest (RF), original param.+NN ensemble of unified RNN-CNN models (DL for deep learning with the default attention), and re-trained RF or transfer DL using incremental amounts of labeled data in each set.}}
\caption{Predicted \red{p$K_i$} values and target specificity for compound DX-9065a interacting with human factor Xa and thrombin.}
\caption{\red{Predicted pIC$_{50}$ values and target specificity for three NSAIDs (CEL: celecoxib, IBU: ibuprofen and ROF: rofecoxib) interacting with human COX-1 and COX-2.}}
\caption{Predicted \red{p$K_i$} values and target specificity for three PTP1B-selective compounds interacting with five proteins in the human PTP family.}
\caption{\red{Interpreting deep learning models for factor Xa binding-site prediction based on joint attention: 3D structure of factor Xa (colored cartoons including helices, sheets, and coils) in complex with DX-9065a (black sticks) (PDB ID:1FAX) where protein SSEs are color-coded by attention scores ($\beta_i$) where warmer colors indicate higher attentions.}}
\caption{\red{Interpreting deep learning models for factor Xa specificity based on joint attentions. Pairwise alignment of amino-acid sequences of factor Xa and thrombin decomposed both sequences into 50 segments (labeled by indices). These segments are scored by one less the average of the corrected attention rank ratios for the two compound-protein interactions. The ground truth of specificity origin is in red.}}
\caption{\red{Comparing the auto-encoding performances between amino acid and SPS sequences using the best seq2seq model (bidirectional GRUs with attention mechanism).}}
\caption{\red{Comparing unified RNN-CNN (SMILES strings for compound representation) and unified RNN/GCNN-CNN (graphs for compound representation) based on RMSE (and Pearson's correlation coefficient) for pIC$_{50}$ prediction.}}
\caption{ Trends of normalised boundary-layer thicknesses appear to scale with the $-1/2$-power law of a wind-based Reynolds number. The boundary layer thicknesses are defined as: the distances to the intercepts (figure \ref{fig:definition_deltau_deltaT}), $\delta_u/H$ ($\circ$) and $\delta_T/H$ ({\tiny$\square$}); the crossovers of dissipation profiles, $\delta^d_u/H$ ({\color{red}$\circ$}) and $\delta^d_T/H$ ({\color{red}\tiny$\square$}); and the displacement thickness, $\delta^*/H$ ({\scriptsize$\times$}). Shown are the Prandtl--Blasius--Pohlhausen $-1/2$-power scaling predictions for vertical natural convection (\ref{eqn:AnalyticalPBVNC}) for $\delta_T$ (\hbox{\drawline{4}{0.5}\spacce{2}\drawline{4}{0.5}\spacce{2}\drawline{4}{0.5}\spacce{2}\drawline{2}{0.5}}) and $\delta_u$ (\hbox{\drawline{1.5}{0.5}\spacce{1.5}\drawline{1.5}{0.5}\spacce{1.5}\drawline{1.5}{0.5}\spacce{1.5} \drawline{1.5}{0.5}\spacce{1.5}\drawline{1.5}{0.5}\spacce{1.5}\drawline{1.5}{0.5}\spacce{1.5}\drawline{1.5}{0.5}\spacce{0.5}}). As reference, the laminar-to-turbulent transition of the shear boundary layer is expected to occur at $(\Rey_{\delta^*})_\textit{cr} \approx 420$ (\hbox{\drawline{4}{0.5}\spacce{1.5}\drawline{1}{0.5}\spacce{1.5} \drawline{4}{0.5}\spacce{1.5}\drawline{1}{0.5}\spacce{1.5}\drawline{4}{0.5}}) \citep{Landau+Lifshitz.1987}. }
\caption{\label{fig:4}(color online). {\bf Unbinding rate of kinesin-1 (experimental data)} (a,b) Distribution-based method: We use the analytical expression in \Eq{eq:pfexp} for a slip bond to fit the empirical cumulative distribution constructed from the experimental data \cite{berger2020}. The numerical values of the fitted parameters determine the unbinding-force pdf $p(F)$ (green line in panel a) and the unbinding rate $\eps(F)$ (red line in panel b). For comparison, the distribution of the experimental data is estimated by a gray histogram. (c,d) Trace-based method: We use all 682 force traces \textemdash one example shown in (c) \textemdash to obtain the unbinding rate $\eps(F)$ as the red line in (d) from \Eq{eq:eps}. Fitting this trace-based estimate with an exponential function (blue line), we obtain the force-free unbinding rate $\eps_0 \simeq 1.1 \, \is$ and the detachment force $\Fd \simeq 7.4 \, \pN$. In (b), the gray lines illustrate the variability of the unbinding rate from bootstrapping; in (d) the errors are given as 95\% confidence intervals \cite{berger2020}.{\color{red}}}
\caption{ Phonon band structures of (a) pristine (Mo$_{2/3}$Y$_{1/3}$)$_2$C, (b) (Mo$_{2/3}$Y$_{1/3}$)$_2$CF$_2$, (c) (Mo$_{2/3}$Y$_{1/3}$)$_2$C(OH)$_2$, and (d) (Mo$_{2/3}$Y$_{1/3}$)$_2$CO$_2$ iMXenes. $\Gamma(0, 0, 0)$, X$(0.5, 0, 0)$, S$(0.5, 0.5, 0)$, and Y$(0, 0.5, 0)$ are high symmetric points of the Brillouin zone. {\color{red}} }
\caption{ Standard deviation of normalized intergranular normal stress (inset: macroscopic tensile stress) as a function of applied strain rescaled by elastic strain calculated with Voronoi finite element simulations assuming crystal elasticity and plasticity for {\red (a)} Al and FCC with various hardening models under uniaxial loading conditions (see the text for models definitions and material parameters used in simulations) {\red and (b) Zn and HCP2 with ideal ($H=0$) plasticity but with different ratios of initial critical resolved shear stresses $\tau_0^{pyr}/\tau_0^{bas}$ (assuming $\tau_0^{pri}=\tau_0^{pyr}$ and $\tau_0^{bas}=100$ MPa).} The fitted logarithmic curves for FCC and HCP2 from Fig. \ref{fig:8c}(a) are shown for comparison.}
\caption{\red Crack length per unit area for different irradiation levels/applied strain \citep{Zhou}. Numerical predictions have been computed assuming an applied true stress of [$455 - 515$] MPa and [$555 - 630$] MPa at the unirradiated state (denoted 0 dpa) (corresponding to the stress range for conventional strain of 17\% and 29\% on the tensile curves given in \citep{Zhou}). For the irradiated materials (2/4/7 dpa), the applied true stress is set equal to the yield stress computed as the sum of the yield stress of the unirradiated material at test temperature ($\sim$180 MPa) and the irradiation hardening reported in \citep{Zhou} (as low strain-hardening is expected for such material). This leads to applied true stress of [$530 - 610$] MPa and [$630 - 710$] MPa for the 2/4 dpa and 7 dpa irradiation levels. The grain boundary length per unit area is approximated by $G_f \approx 2/\phi$, where $\phi$ is the grain size. The critical stress $\sigma_c$ was calibrated to a value of 550 MPa.}
\caption{The latent space of the deterministic autoencoder (DAE), variational autoencoder (VAE), Wasserstein autoencoder (WAE), as well as WAE with by our KL penalty. {\color{blue}Blue} circles: Posterior or aggregated posterior distributions of data in the latent space. {\color{red}Red} circles: Regularizations of the posterior.}
\caption{ error ({\color{blue}-}) and control line ({\color{red}-.-})}
\caption{Result of motion capture completion (motion interpolation) of a walking sequence in a Human 3.6M dataset. Reconstructed poses are shown together with the original poses for a qualitative comparison. An attention matrix of all time steps, together with detailed attention weights of time step $t=75$ are shown in \textcolor{color_dproposed}{\rule[0.06cm]{0.4cm}{1.4pt}}. To reconstruct the pose of $t=75$, the network combines information from other poses with the similar appearance, $t=114, 111, 35$. Attention weights of previous method~\cite{conf:cln:SOnderby:2015} are also given in \textcolor{color_battention}{\rule[0.06cm]{0.4cm}{1.4pt}} for comparison. }
\caption{Simulation results: often-used multi-loop SISO ILC design (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[black,thick][-] (0,-3) -- (6,3);\protect\draw[black,thick][-] (0,3) -- (6,-3);}) can lead to non-convergent algorithms. Through the developed design approaches in steps {\protect\circled{4}} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[red,thick] (0,-3) rectangle ++(6,6);}), {\protect\circled{5}} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[blue,thick] (7,0) circle (3);}), and {\protect\circled{6}} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[green,thick] (0,0) +(3,0) -- +(0,3) -- +(-3,0) -- +(0,-3) -- cycle;}), a well-motivated balance can be made between achievable performance, i.e., norm of the asymptotic error $\|e_\infty\|_F$, and the associated user effort in terms of modeling cost and design complexity.}
\caption{Oc\'e Arizona 550GT flatbed printer. The carriage moves along the gantry, which provides the motion freedom to cover the printing surface. The actuator forces are indicated by red arrows. The inputs considered for control are$F_L$, $F_R$, and the outputs are $x_L$, $\varphi_2$, indicated by blue arrows.}
\caption{Bode diagram of non-decoupled true plant $G_o(z)$ (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[gray,thick][-] (0,0) -- (10,0);}), true plant $G$ (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[blue,thick][-] (0,0) -- (10,0);}) after decoupling transformations in step \protect\circled{3} of \procref{MIMO_ILC:proc_MIMO_ILC}, and model of decoupled plant $\widehat{G}(z)$ (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[red,thick][-] (0,0) -- (4,0);\protect\draw[red,thick][-] (6,0) -- (10,0);}) used for ILC design.}
\caption{Reference trajectories $r_x$ (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[black,thick][-] (0,0) -- (10,0);}) and $r_\varphi$ (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[gray,thick][-] (0,0) -- (4,0);\protect\draw[gray,thick][-] (6,0) -- (10,0);}). The start of the motion tasks are indicated by dotted lines.}
\caption{Error signals in trial $j=10$ of robust SISO design in step \protect\circled{4} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[red,thick][-] (0,0) -- (10,0);}), decentralized design in step \protect\circled{5} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[blue,thick][-] (0,0) -- (10,0);}) and centralized design in step \protect\circled{6} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[green,thick][-] (0,0) -- (10,0);}).}
\caption{Feedforward signals in trial $j=10$ of the robust SISO design in step \protect\circled{4} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[red,thick][-] (0,0) -- (10,0);}), decentralized design in step \protect\circled{5} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[blue,thick][-] (0,0) -- (10,0);}) and centralized design in step \protect\circled{6} (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{ \protect\draw[green,thick][-] (0,0) -- (10,0);}). The start of the motion tasks are indicated by dotted lines.}
\caption{\footnotesize{\textbf{Optical properties of halide perovskites.}} \red{(a) Photoluminescence peak energy and FWHM for pure chloride, bromide, and iodide perovskite thin films in the cubic phase.} \red{(b)} Refractive indices $n$ and \red{(c)} extinction coefficients $k$ of three commonly used perovskite composites: MAPbCl$_3$, MAPbBr$_3$, and MAPbI$_3$, shown together with the corresponding results for SiO$_2$, Si, and GaAs, for comparison. The data for the perovskites are obtained by Leguy {\it et al.} through single crystal ellipsometry and from Refs.~\cite{rodriguez2016self,schinke2015uncertainty,ozaki1995spectroscopic} for SiO$_2$, Si, and GaAs. These data are collected from Refs.~\cite{wang2016visible, kim2017high, ko201515, imran2018benzoyl, slimi2017synthesis, yuan2015inverted, kondo2005room, sebastian2015excitonic, wang2017stabilizing}.}
\caption{\textbf{Harmonics generation from halide perovskites.} Spectra of photoluminescence (PL), second-harmonic (SHG) and third-harmonic generation (THG) from CsPbBr$_3$ at different excitation wavelengths (of a femtosecond laser): (a) 1700~nm and (b) 1600~nm. \red{Insets schematically show level diagrams and processes of SHG, THG, and PL at corresponding pulsed laser optical excitation. Adapted with permission from~\cite{clark2016polarization}. Copyright 2018 by the American Physical Society.}}
\caption{\textbf{Resonant halide-perovskite nanoparticles.} (a) \red{Analytically calculated total scattering efficiency $Q_{scat}$ of a spherical MAPbBr$_3$ nanoparticle in vacuum. Overlaid lines depict approximate dispersion of the electric dipole, magnetic dipole, and magnetic quadrupole resonances of the particle. The arrow indicates position of MAPbBr$_3$ exciton.} (b) Calculated near-field distributions for \red{single MAPbI$_3$ nanoparticles supporting the magnetic dipolar and quadrupolar resonances at emission wavelength}.\red{Adapted with permission from~\cite{tiguntseva2018light}. Copyright 2018 American Chemical Society.} (c) PL enhancement for a nanoparticle at magnetic quadrupole resonance \red{(red curve)}, as compared with thin a film \red{(black curve)} and non-resonant nanoparticle \red{(green curve)}. \red{Adapted with permission from~\cite{tiguntseva2018light}. Copyright 2018 American Chemical Society.} (d) Lasing modes in CVD-grown CsPbBr$_3$ submicron particles of different sizes.\red{Reprinted with permission from~\cite{tang2017single}. Copyright 2018 American Chemical Society.} (e) Lasing from perovskite submicron \red{CsPbX$_3$ spheres with different halogens X: Cl,Br, I and mixed ones}. \red{Reprinted with permission from~\cite{tang2017single}. Copyright 2018 American Chemical Society.}}
\caption{\textbf{Strong coupling of excitons to Fabry-Perot modes in perovskite nanowires.} (a) Schematic of a perovskite nanowire photoexcitation and detection. Inset: electron image of the MAPbBr$_3$ nanowire. (b) Calculated electric near-field distribution in the nanowire placed on glass and silver substrates. (c) Dispersion curves for exciton-polaritons in the nanowire on a silver substrate with different SiO$_2$ spacers (5, 10, 20~nm) and without silver. Solid lines represent the fittings by formulas similar to Eq. (10). \red{Adapted with permission from~\cite{shang2018surface}. Copyright 2018 American Chemical Society.}}
\caption{\textbf{Perovskite metasurfaces.} \red{Simulated (a) electric and (b) magnetic field distributions for TE polarized incident light of metasurfaces. (c) Unpolarized optical microscope image of nanograting tunable color perovskite metasurfaces 350~nm with different mill depths ranging from $\approx$200~nm to $\approx$20~nm. Scale bar indicates 20~$\mu$m. (d) CIE color palette with marked points for a selection of the nanograting metasurfaces~\cite{gholipour17}.}}
\caption{\textbf{Integration of halide perovskites with non-perovskite nanostructures.} (a) An embedded dielectric/metal heterostructure for plasmonic output of dielectric laser. Left: schematic illustration. Central: \red{electron} image in false colors. Right: optical images in reflection and luminescent modes~\cite{li2016output}. (b) Left: \red{electron} image of the MAPbBr$_3$ microwire on silicon grating, where the scale bar is 10 $\mu$m. Right: fluorescent microscope image \cite{wang2016high}. (c) Schematic showing the time-delayed pump excitations at 800 and 400~nm, which photogenerate free carriers in MAPbI$_3$ and MAPbBr$_3$, respectively~\cite{chanana2018ultrafast}.}
\caption{\textbf{Example Video Segmentations:} We show results from our self-supervised model on the task of video segmentation. Colors indicate different instances. Although the model is trained without ground truth labels, the model can still propagate segmentations throughout videos. The left column shows the input frame and input masks to the model, and the rest show the predictions. Results suggest that the model is generally robust to intra-class variations, such as deformations, and occlusions. The model often handles multiple objects and cluttered backgrounds. Best viewed in color. We provide videos of results online at \textcolor{blue}{\url{https://goo.gl/qjHyPK}}}
\caption{\label{fig:ev-Method} \blue{Test-Method-level test prioritization in evolution}}
\caption{\blue{TCP results on geojson-jackson (APFD)}}
\caption{\blue{TCP results on geojson-jackson (APFDc)}}
\caption{\blue{TCP results on javapoet (APFD)}}
\caption{\blue{TCP results on javapoet (APFDc)}}
\caption{\label{fig:ev-Class} \blue{Test-Class-level test prioritization in evolution}}
\caption{\blue{TCP results on geojson-jackson (APFD)}}
\caption{\blue{TCP results on geojson-jackson (APFDc)}}
\caption{\blue{TCP results on javapoet (APFD)}}
\caption{\blue{TCP results on javapoet (APFDc)}}
\caption{(a,b) Relaxation times $\tau^j_\mathrm{LO}$ and $\tau_\mathrm{log}$ obtained by fitting Eq. \ref{stretch_exp_fit} or \ref{log_fit} to $f^{j}(q,\Delta t)$, obtained from simulations for $\phi = 0.62$, $x_\mathrm{s} = 0.01$ ($\Box$), 0.05 (\textcolor{red}{\Large $\circ$}), 0.10 (\textcolor{blue}{$\triangle$}) and (a) $\delta = 0.20$, (b) $\delta = 0.35$. Full symbols: Large particles; Open Symbols: Small particles. (c) Relaxation times $\tau^S$ and $\tau_\mathrm{log}$ for the small particles from experiments, for $\phi =0.61$, $x_\mathrm{s} = 0.01$ and $\delta = 0.18$ ($\Box$) and $\delta = 0.28$ ($\triangle$). Open symbols: Long-time relaxation times $\tau^S_\mathrm{LO}$. Full symbols: Short time relaxation times $\tau^S_\mathrm{SH}$.}
\caption{(a,b) Plateau heights $f_c$ obtained by fitting Eq. \ref{stretch_exp_fit} or \ref{log_fit} to $f(q,\Delta t)$ from simulations for $\phi = 0.62$, $x_\mathrm{s} = 0.01$ ($\Box$), 0.05 (\textcolor{red}{\Large $\circ$}), 0.10 (\textcolor{blue}{$\triangle$}) and (a) $\delta = 0.20$, (b) $\delta = 0.35$. Full symbols: Large particles; Open symbols: Small particles. (c) Plateau heights $f_c$ for the small particles from experiments for $\phi =0.61$, $x_\mathrm{s} = 0.01$ and $\delta = 0.18$ ($\Box$) and $\delta = 0.28$ ($\triangle$).}
\caption{Parameters obtained by fitting Eq. 6 of the main manuscript to $f(q,t)$ from simulations for $\delta = 0.2$, $\phi = 0.60$ (\textcolor{green}{$\triangle$}), 0.61 (\textcolor{blue}{$\triangledown$}), 0.62 (\textcolor{red}{$\Box$}) and 0.63 ({\Large $\circ$}) and $x_\mathrm{s} = 0.01$ (left column), 0.05 (mid column) and 0.1 (right column), as a function of $q\sigma_L$. (a), (e), (i): Long-time relaxation times for the small particles, $\tau^S_\mathrm{LO}$ (open symbols). (b), (f), (l): Long-time relaxation times for the large particles, $\tau^L_\mathrm{LO}$ (full symbols). (c), (g), (m): Plateau height for the large, $f_{c}^{L}$ (full symbols, when present), and small, $f_{c}^{S}$ (open symbols), particles. (d), (h), (n): Stretching exponent $\beta^S$ obtained from the fits to the long-time relaxation of the small particles.}
\caption{Parameters obtained by fitting Eq. 6 of the main manuscript to $f(q,t)$ from simulations for $\delta = 0.35$, $\phi = 0.62$ and $x_\mathrm{s} = 0.01$ ($\Box$), 0.05 (\textcolor{red}{\Large $\circ$}), 0.10 (\textcolor{blue}{$\triangle$}), as a function of $q\sigma_L$. (a), (e), (i): Long-time relaxation times for the small particles, $\tau^S_\mathrm{LO}$ (open symbols). (b), (f), (l): Long-time relaxation times for the large particles, $\tau^L_\mathrm{LO}$ (full symbols). (c), (g), (m): Plateau height for the large, $f_{c}^{L}$ (full symbols, when present), and small, $f_{c}^{S}$ (open symbols), particles. For $x_\mathrm{s} = 0.01$ we report the logarithmic non-ergodicity parameter $f_\mathrm{log}^S$ as defined in Eq.7 of the main manuscript. (d), (h), (n): Stretching exponent $\beta^S$ obtained from the fits to the long-time relaxation of the small particles.}
\caption{Parameters obtained by fitting Eq.6 to $f(q,t)$ from experiments for $x_\mathrm{s} = 0.01$, $\delta = 0.18$ and $\phi = 0.60$ (\textcolor{green}{$\triangle$}), 0.61 (\textcolor{blue}{$\triangledown$}), 0.625 (\textcolor{red}{$\Box$}) (Left) and $\delta = 0.28$ $\phi = 0.58$ (\textcolor{green}{$\triangle$}), 0.59 (\textcolor{blue}{$\triangledown$}), 0.60 (\textcolor{red}{$\Box$}) 0.61 ({\Large $\circ$}) (Right). (a) ,(d) Long and short relaxation times of the small particles, $\tau_\mathrm{LO}$ (open symbols) and $\tau_\mathrm{SH}$ (full symbols), respectively. (b), (e): Plateau height $f_c^S$ (b) and $f_\mathrm{log}^S$ (e), where the last is defined in Eq.7 of the main manuscript. (c), (f) Stretching exponents obtained $\beta^S$ obtained from the fits to the long-time relaxation of the small particles.}
\caption{\label{fig:orb} Graphical comparison of the key physical quantities in regimes I and II. (a) Mean value of the orbit radius, $\langle r \rangle$, as a function of optical power. The insets show the probability density of the particle positions in the lateral $xy$ plane relative to the given scale bar. Data related to circularly polarized beams (CP) are denoted by {$\bigcirc$}, $\bigtriangleup$ and $\bigtriangledown$ for the orbiting regime II. $\bigtriangleup$ and $\bigtriangledown$ symbols correspond to increasing and decreasing optical power, respectively. For comparison data corresponding to parallel linear polarization (LP) are marked by black $+$ signs. The sequence in which the data was acquired is indicated by colors and markers {\color{red}{$\bigcirc$}}, {\color{red}{$\bigtriangledown$}}, {\color{blue}{$\bigtriangleup$}}, {\color{blue}{$\bigcirc$}}, {\color{black}{$+$}}, and {\color{green}{$\bigcirc$}}. Overlapping points {\color{red}{$\bigcirc$}} with {\color{red}{$\bigtriangledown$}} and {\color{blue}{$\bigtriangleup$}} with {\color{blue}{$\bigcirc$}} illustrate how different calculation methods, used for regime I and II, overlap in the region where the orbiting is not fully developed. (b) Resonant trap frequency $\omega_0/2 \pi$ or orbiting frequency $\Omega/2 \pi$ for regime I or II, respectively. The theoretical value from Eq. (\ref{eq:oeqm_frq}), is indicated by the horizontal red line. Measured values of $\xi/m$, and calculated values of $f_r/f_{\phi}$ have been used. (d) Spin force $f_{\phi}$ for a particle of radius $770$ nm and density 2200 kg/m$^3$. The presented values were determined as the mean value for $x$ and $y$ directions from fits to the PSD (Supplementary Information ) for regime I and from the orbiting equation (Eq. \ref{eq:oeqm_frq}) for regime II (see details in Supplementary Information). The grey curve shows the theoretical force, obtained from generalized Lorentz Mie theory (Supplementary Information). A single scaling parameter has been used (Supplementary Information ). }
\caption{(Color online) Systematics along O, Ca and Ni isotopic chains: (a) absolute binding energy, (b) two-neutron separation energy, (c) neutron-number dispersion, (d) perturbative correction to the average neutron number. Plot markers correspond to HFB (\bluecircle), second-order BMBPT (\redsquare) and third-order BMBPT (\orangediamond). Experimental values are shown as black bars \cite{WaAu12}. }
\caption{Absolute ground-state binding energies (top) and two-neutron separation energies (bottom) along O, Ca and Ni isotopic chains. Results are displayed for second-order BMBPT (\redsquare), second-order NCSM-PT (\bluecircle), large-scale IT-NCSM (\orangestar), GSCGF-ADC(2) (\greentriangleup), MR-IMSRG(2) (\hspace{1.7pt}\purplepentagon) and CR-CC(2,3) (\orangediamond). Experimental value are shown as black bars \cite{WaAu12}. }
\caption{Computational runtime versus mass number from BMBPT(2) (\redsquare), BMBPT($3^*$) (\orangediamond), MR-IMSRG(2) (\hspace{2pt}\purplepentagon) and ADC(2) calculations. }
\caption{Profile of y for GM1 {\color{red}}}
\caption{When two doubly degenerate bands with positive parity eigenvalues and two doubly degenerate bands with negative parity eigenvalues are inverted at a TRIM {\blue point~\cite{HigherOrderTIBernevig,HOTIBismuth}, the occupied bands cannot be expressed as a linear combination of EBRs~\cite{ZakBandrep1,ZakBandrep2,QuantumChemistry,Bandrep1,Bandrep2,Bandrep3,JenFragile1}} and the $\mathbb{Z}_{4}$ topological index~\cite{ChenTCI,AshvinIndicators,AshvinTCI,EslamInversion} is changed by 2. In a $\mathcal{T}$-symmetric crystal with vanishing SOC, this process may nucleate a pair of Dirac nodal lines with nontrivial monopole charge (MNLs)~\cite{YoungkukMonopole}{\red (dashed lines in left panel)}. On the 1D hinges of this system, the projections of the MNLs will be spanned by nearly-flat hinge states (an explicit model is provided {\red in SM~\ref{sec:TBmodel})}, which represent the $d-2$-dimensional generalization of drumhead surface states. These hinge bands present an example of higher-order topology in a bulk-gapless system, and are the spinless analogs of the quadrupolar hinge bands predicted in certain tetragonal Dirac semimetals~\cite{HingeSM,TaylorToy}. When $\mathcal{I}$-symmetric SOC is introduced, the system will necessarily gap into a HOTI if all other bands are uninverted, and the flat-band hinge states will open into helical pairs spanning the bulk and surface gaps. HOTIs driven by this ``double band inversion'' include bismuth~\cite{HOTIBismuth} and $\beta$-MoTe$_{2}$ (Fig.~\ref{fig:Fig2}(d)).}
\caption{(a) The monoclinic lattice of $\beta$- (1T'-) MoTe$_2$~\cite{XTe2Structures} in SG 11 $P2_{1}/m$. (b,d) Bulk bands of $\beta$-MoTe$_2$ calculated without and with the effects of SOC incorporated, respectively. Double band inversion occurs about the $\Gamma$ point, as indicated by the parity eigenvalues in (d). When SOC is neglected, a time-reversed pair of nodal lines forms, intersecting $Y\Gamma$ in (b) {\blue connecting the 28$^\text{th}$ and 29$^\text{th}$ spin-degenerate pair of bands (red dots in (b)). (c) These nodal lines are irregularly shaped, and lie in time-reversed pairs between $k_{y}= \pm 0.06\ (2\pi/b)$ and $k_{y}= \pm 0.19\ (2\pi /b)$. We surround one of these nodal lines with a closed, tetragonal prism and calculate the Wilson loop {\red around $k_{z}$-normal planes} as a function of the azimuthal momentum $k_{\theta}$ (exact coordinates provided {\red in SM~\ref{sec:DFTmonopole})}; the results are shown in the inset panel in (b).} $\mathbb{Z}_{2}$-nontrivial winding is clearly visible, and indicates a nontrivial monopole charge~\cite{YoungkukMonopole,AdrianMonopole,SigristMonopole}. (d) When SOC is introduced, a direct gap develops at all crystal momenta (gray shaded region), realizing a band gap at the Fermi energy with the $\mathbb{Z}_{4}$ parity index (Table~\ref{tb:z4}) of a HOTI~\cite{ChenTCI,AshvinIndicators,AshvinTCI,EslamInversion}. The details of our first-principles calculations can be {\red found in SM~\ref{sec:DFT}.}}
\caption{The number of Kramers pairs with $-1$ parity eigenvalues at each of the TRIM points in $\beta$-MoTe$_2$, obtained from {\red first-principles (SM~\ref{sec:DFTmethod})}. The $\mathbb{Z}_{4}$ index~\cite{ChenTCI,AshvinIndicators,AshvinTCI,EslamInversion}, calculated by their sum modulo $4$, is $2$, indicating that $\beta$-MoTe$_2$ is a HOTI.}
\caption{{\blue (a) Bulk Wilson loops calculated for $\beta$-MoTe$_2$ from first principles in the absence of {\red SOC (SM~\ref{sec:DFTnested})}. For all values of $k_{y}$ away from the MNLs in Fig.~\ref{fig:Fig2}, the $z$-directed Wilson loop spectrum $W_{1}(k_{x},k_{y})$ exhibits a large Wilson gap at $\theta_{1}/2\pi \approx \pm 0.25$; representative examples are shown in (c,d) for $k_{y} = 0,\pi$, respectively. (b)} Calculating the determinant of the \emph{nested} Wilson matrix $W_{2}(k_{y})$ of the Wilson bands within this gap~\cite{multipole,WladPump,HigherOrderTIBernevig,HingeSM}, we {\blue find that it is quantized at $\pm 1$ by the combined antiunitary symmetry $(\mathcal{I}\times\tilde{\mathcal{T}})^{2}=+1$ {\red (SM~\ref{sec:DFTnested})}, and jumps as it passes over a bulk MNL, indicating that the planes above and below the MNL} are topologically distinct~\cite{multipole,WladPump,HigherOrderTIBernevig}.}
\caption{{\blue (a-c) Schematic surface state evolution of a HOTI driven by double band inversion. (a) Two bulk bands inverted at the same energy (blue dashed lines) realize a fourfold surface Dirac fermion (purple lines)~\cite{HOTIBismuth}. (b) In the absence of multiple surface glide reflection symmetries, this fermion is unstable~\cite{DiracInsulator}, and will split into two, twofold surface fermions, which may be stabilized by either a surface mirror (topological crystalline insulator)~\cite{TeoFuKaneTCI,HsiehTCI} or glide symmetry (hourglass insulator)~\cite{HourglassInsulator,DiracInsulator}. (c) In the absence of surface reflection symmetries, these twofold cones (yellow circles in (b), dashed lines in (c)) hybridize and gap, realizing the surface of a HOTI~\cite{HigherOrderTIBernevig}.} (d) Spectral weight at the Fermi energy of states on the (001) surface of $\beta$-MoTe$_2$, calculated from first {\red principles (SM~\ref{sec:DFTmethod})}, and plotted as a function of the in-plane momenta $k_{x,y}$, and {\blue (e)} along $k_{x}=0$ as a function of energy. Each of the two band inversions at the bulk $\Gamma$ point (Fig.~\ref{fig:Fig2}(d)) nucleates a topological twofold surface {\blue cone centered at $k_{x}=k_{y}=0$ (purple); {\red the cones then} repel each other in energy and merge with the projections of the bulk states (b,e). As schemicatically depicted in (c), the surface bands from these cones {\red (red arrows in (d))} hybridize and gap (yellow dashed lines in (e))} to form a narrowly avoided crossing. {\blue In $\gamma$-XTe$_{2}$, these hybridized cones also appear as surface states~\cite{wang_mote2:_2016,soluyanov_type-ii_2015,sun_prediction_2015,WTe2Arpes1,WTe2Arpes2,WTe2Arpes3,WTe2Arpes4,MoTe2Arpes1,MoTe2Arpes2,MoTe2Arpes3,MoTe2Arpes4,WTe2STM,MoTe2STM}, but their gap is spanned by small, topological Fermi arcs from bulk type-II Weyl points.}}
\caption{(a) Primitive orthorhombic BZ~\cite{BCTBZ}. Bulk bands for $\tilde{\mathcal{H}}(\vec{k})$ (Eq.~(\ref{eq:simpleHam})) with the parameters in Eq.~(\ref{eq:uncoupledParams}). In this limit, this eight-band model exhibits an overall $SU(2)$ spin rotation symmetry, {\blue and bands therefore appear in spin-degenerate pairs. In the inset panel, we highlight a} monopole nodal line (MNL) at $E_{F}=0$ along $\Gamma Z$ {\blue (red circles)} that is linked to its time-reversal partner by nodal lines encircling the $\Gamma$ {\blue point} directly above and below $E_{F}$ {\blue (blue circle) ({\red axes not in parenthesis in} Fig.~\ref{fig:coupled}(d))}, as discussed in Ref.~\onlinecite{YoungkukMonopole}. {\red Using the bulk tight-binding model, we deduce that the MNLs lie in the $k_{yz}$-plane.} (c) The Wilson loop eigenvalues over the lower four occupied bands on a sphere surrounding this nodal line, plotted as a function of the azimuthal momentum $k_{\theta}$, exhibit helical winding. As detailed in Refs.~\onlinecite{YoungkukMonopole,AdrianMonopole,SigristMonopole}, this confirms that this nodal line carries a nontrivial monopole charge, {\red and thus, is a MNL}. (d,e) The $(100)$ and $(010)$ surface states of this model, respectively, {\blue calculated at E=0}. {\red In (d), we observe both drumhead states on the interior projections of the bulk MNLs (white arrows)~\cite{YoungkukLineNode} and extraneous surface states (red arrows) that are remnants of the bulk double band inversion. Specifically, the interior states indicated by the white arrows are topologically protected~\cite{YoungkukMonopole}, whereas the line of states shown by the red arrows is topologically trivial and lies outside the projections of the MNLs. These states originate from the double band inversion shown in Fig.~\ref{fig:DBI}: the first band inversion created a drumhead state at the origin of the surface BZ, and the pinching process of the line node at half filling formed a second set of surface states, rather than removed the drumhead states from the first band inversion. These extraneous surface states are topologically trivial, and represent an artificial degeneracy in this limit of this model; we add the term necessary to hybridize and gap then in Fig.~\ref{fig:coupled}.} (f) The bands of an $x$-directed slab of this model, plotted at $k_{y}=0$ as a function of $k_{z}$. {\red The extraneous surface states, in (d) are marked here with a red arrow as well.}}
\caption{{\red Bulk bands of $\tilde{\mathcal{H}}_{C}(\vec{k})$ in Eq.~(\ref{eq:coupledHam}), plotted along $\Gamma Z$ in the vicinity of the $\Gamma$ point with the parameters in Eqs.~(\ref{eq:uncoupledParams}) and tuning $v_{z}$ between $0$ and $2$. Throughout this process, bands become doubly inverted, and eventually form a time-reversed pair of MNLs at half-filling (red circles at $v_{z}=2.0$), as described in Ref.~\onlinecite{YoungkukMonopole}}.}
\caption{{\red (a,b) $x$-directed Wilson loops evaluated for the lowest two spinless pairs of bands in Fig.~\ref{fig:uncoupled}(a) at $k_{z} = 0,\pi$, respectively. The winding of the Wilson loop at $k_{z}=0$ is not protected by spinful time-reversal symmetry, as it would be in a 2D topological insulator~\cite{AndreiTI}, but is instead protected by the combination of the bulk inversion eigenvalues~\cite{Alexandradinata14} and the absence of additional bands in the Wilson projector. {\blue At $k_{z}=\pi$ (b), the bulk inversion eigenvalues require that the $\theta = 0$ at $k_{y}=0,\pi$~\cite{Alexandradinata14}; the $x$-directed Wilson loop in this plane exhibits trivial winding, and extremely weak dispersion.}}}
\caption{(a) Bulk bands for $\tilde{\mathcal{H}}_{C}(\vec{k})$ in Eq.~(\ref{eq:coupledHam}), plotted with the parameters in Eqs.~(\ref{eq:uncoupledParams}) and~(\ref{eq:coupledParams}). (b,c) $(100)$ and $(010)$ surface Green's function at $E_{F}=0$ for the same model, respectively. {\red Topological} drumhead states appear on both surfaces {\red in the interior projections of the MNLs (white arrows)}, indicating that the MNLs have {\blue become tilted} and now have nonzero projections in both the $x$ and $y$ directions. {\blue Crucially, the extraneous} surface spectral weight spanning the projections of the MNLs from Fig.~\ref{fig:uncoupled}(d) has been lifted. {\blue (d) Specifically, we observe that both the nodal lines at half-filling (red), and the large nodal line directly below it in energy (blue) have become tilted by $\sim 45^\circ$ about the $k_{z}$-axis (axes in parenthesis are those after including Eq.~(\ref{eq:coupledHam})). (e)} The bands of an $x$-directed slab of this model, plotted at $k_{y}=0$ as a function of $k_{z}$, confirm that the {\blue extra surface states} have hybridized and split. {\blue (f)} The bands of a $z$-directed rod of this model (finite in the $x$ and $y$ directions). Flat-band-like 1D states can be observed spanning the hinge projections of the MNLs.}
\caption{(a) Bulk and (b) hinge bands of a $z$-directed rod of $\tilde{\mathcal{H}}_{C}(\vec{k})$ (Eq.~(\ref{eq:coupledHam})) with the Kane-Mele-like SOC term $V_{HOTI}(\vec{k})$ (Eq.~(\ref{eq:HOTI}), plotted with the parameters in Eqs.~(\ref{eq:uncoupledParams}) and~(\ref{eq:coupledParams}) and $v_{H}=1.2$. The flat-band hinge states from Fig.~\ref{fig:coupled}(e) have evolved into a pair of 1D helical modes, confirming that $V_{H}(\vec{k})$ induces a phase transition from a monopole NLSM to a $\mathcal{T}$-symmetric HOTI. When $V_{Axion}(\vec{k})$ (Eq.~(\ref{eq:axion})) is used instead of $V_{HOTI}(\vec{k}))$, the bulk bands appear similar to those in (a), but the hinge spectrum instead exhibits oppositely propagating spin-degenerate pairs of chiral modes on {\blue opposing} hinges.}
\caption{ Demonstration videos for reacher task. Sample video from training set \textbf{(top)} and meta-test \textbf{(bottom)}. Each of the video contains a video reaching to two multiple targets; (\textcolor{orange}{orange}, \textcolor{green}{green}) for training set, and (\textcolor{SpringGreen}{light green}, \textcolor{Fuchsia}{purple}) for meta-test set. Rest two colors are distractors. }
\caption{Steps in generating a dynamic abdominal ECG mixture as implemented in the \textit{fecgsyn} toolbox \highlightcolor{(a)} Vector loop representing the time-varying source model \highlightcolor{(b)} Homogeneous volume conductor model representing the maternal-fetal anatomy \highlightcolor{(c)} Resulting surface potentials where numbered squares indicate sensor positions. Note the P wave, QRS complex and T wave are oriented differently in 3D space.}
\caption{32 weeks GA maternal-fetal model with fetus in right occiput transverse (ROT) presentation \highlightcolor{(a)} Abdominal view \highlightcolor{(b)} Sagittal view \highlightcolor{(c)} Coronal view where \protect\mycbox{shadedFetus} = fetus, \protect\mycbox{shadedAmniotic} = amniotic fluid and \protect\mycbox{shadedAbdomen} = maternal body}
\caption{\highlightcolor{(a)} Fetal body maps\protect\footnotemark\:where A = Arms, B = Back, C = Chest, H = Head, I = Inguinal, L = Legs, S = Buttocks.\highlightcolor{(b)} Distribution of vernix caseosa in each region as identified from fetal body maps. Total cases (n=100) where cases $<$37 weeks GA (n=30), 37.1$-$40.9 weeks GA (n=59) and $>=$ 41 weeks GA (n=11). Data from \cite{Archana_ClinicalStudySurface_2008}}
\caption{\highlightcolor{(a)} Cross section (y $>$ 127mm) of tetrahedral finite element model with vernix caseosa generated at 2mm thickness in the head region \mbox{\highlightcolor{(b)} Inset view} showing fine discretization in the vernix caseosa compartment. Tissue types are color coded as follows: \protect\mycbox{fetus} = fetus, \protect\mycbox{vernix} = vernix caseosa, \protect\mycbox{amniotic} = amniotic fluid and \protect\mycbox{abdomen} = maternal body}
\caption{Developed process for generating surface potentials in an asymmetric volume conductor model. Dark gray boxes indicate data from the labeled source. Light gray arrows indicate a processing step using the labeled tool. Black arrows indicate data input to the target. \highlightcolor{(a)} indicates steps utilized for model setup and \highlightcolor{(b)} indicates steps utilized for the experiments.}
\caption{Model refinement process showing lead field matrix convergence for the Back 1mm model as maximum element volume per compartment is halved, approximately doubling the number of tetrahedra at each step. \protect\includegraphics[width=0.02\textwidth]{circ.pdf} shows a converging relative difference measure ($\norm{RDM^*}_2$) calculated at each step compared to the subsequent solution, \protect\includegraphics[width=0.02\textwidth]{tri.pdf} shows time to compute tetrahedralization via TetGen and \protect\includegraphics[width=0.02\textwidth]{square.pdf} shows time to compute lead field matrix via SimBio per sensor. Subsequent reference solution for the final data point is not shown.}
\caption{Cross section (y $>$ 127mm) of selected volume conductor models and resulting surface potentials produced by dipoles along each co-ordinate axis with rows representing \highlightcolor{(a)} Homogeneous, \highlightcolor{(b)} No Vernix, \highlightcolor{(c)} Back 3mm, \highlightcolor{(d)} Chest 3mm and \highlightcolor{(e)} Head 3mm volume conductor models. \protect\myfilledcircle{sensor1} indicates position of the fetal source and tissue types are color coded as follows: \protect\mycbox{fetus} = fetus, \protect\mycbox{vernix} = vernix caseosa, \protect\mycbox{amniotic} = amniotic fluid, \protect\mycbox{abdomen} = maternal body. Note: The color scale is asymmetric around 0 V/Am, indicating maximum negative potentials are approximately twice as large as maximum positive potentials.}
\caption{Simulated NI-FECG waveforms for the homogeneous and asymmetric volume conductor models \highlightcolor{(a)} shows maternal body model with 6 sensor positions (\protect\myfilledcircle{sensor7},\protect\myfilledcircle{sensor5},\protect\myfilledcircle{sensor3},\protect\myfilledcircle{sensor4},\protect\myfilledcircle{sensor2},\protect\myfilledcircle{sensor6}), reference node (\protect\myfilledcircle{sensor8}) and fetal source position (\protect\myfilledcircle{sensor1}) \highlightcolor{(b)} shows potentials observed at each sensor position with respect to the reference node for the homogeneous (\protect\mysolidline{homogeneous}) and asymmetric volume conductor models: No Vernix (\protect\mysolidline{model1}), Back 3mm (\protect\mysolidline{model4}), \mbox{Chest 3mm (\protect\mysolidline{model3})}, Head 3mm (\protect\mysolidline{model2}), BackChest 3mm (\protect\mysolidline{model7}), BackHead 3mm (\protect\mysolidline{model6}), ChestHead 3mm (\protect\mysolidline{model5}), BackChestHead 3mm (\protect\mysolidline{model8}). \protect\mycbox{qrszone} indicates the QRS complex detection zone and \protect\mycbox{tzone} indicates the T wave detection zone.}
\caption{Spectral Energy Distribution (SED) of Vela X. We show separately the spectra for the low-energy (LE) and high-energy (HE) components as derived in this work. The lines and shaded bands show the best-fit power laws for each component with their uncertainties, based on the fit using model A. The darker shaded band corresponds to 68\% statistical uncertainties only, while the lighter shaded band corresponds to the sum in quadrature of statistical uncertainties and systematic uncertainties from the morphological representation of the sources (models B and C) and the LAT effective area (see Section~\ref{sec:spectrum} for details). The points and 95\% confidence level upper limits show the binned SED. For points capped error bars show the 68\% statistical uncertainties only, while the uncapped error bars show the sum in quadrature of statistical uncertainties and systematic uncertainties. Upper limits include systematic uncertainties as well. We also show the overall SEDs of Vela~X from H.E.S.S. \citep{hessvelax2012}, from the LAT measurements $< 100$~GeV \citep{grondin2013}, and from the 3FHL catalog \citep{3FHL}. The dashed lines \highlight{show} the predictions of the radiative model described in Section~\ref{sec:discuss}. }
\caption{(Left) $f^{Ly\alpha}_{esc}$ versus the dust optical depth $\tau_a$ for different geometries in outflows with the same physical properties ($V_{\rm exp}$ and $N_{\rm H}$), as indicated in the figures. The output of the radiative transfer code is represented by \colorThin\circles,\colorWind\diamonds and\colorBicone\squares for the{\it Thin Shell}, galactic wind and biconical geometries respectively. Additionally, our analytical fit is represented by solid lines with the same color code as the code's output. (Right) \lya\line profile for different geometries with the same physical properties. In colored lines the radiative transfer code output is plotted for the{\it Thin Shell} geometry (\colorThin), the galactic wind (\colorWind) and the biconical galactic wind (\colorBicone). }
\caption{ LAE LF at redshift 2.2 (top left), 3.0, (top right), 5.7 (bottom left) and 6.7 (bottom right). The LF computed for different geometries is plotted as colored continuum lines, in \colorWind\for the{\it Wind} geometry, in \colorBicone\for the{\it Bicone} geometry and in \colorThin\the{\it Thin Shell} geometry. In continuum black we show the intrinsic \lya\LF. The black dashed lines show the combined LF that is fitted that, at the same time, is the\NoRT\LF (detailed in\S \ref{ssec:Results}) LF. At redshift 2.2 we also show the LF observed by Kono et al 2016 (blue dots), Sobral et al. 2016 (purple diamonds) and Cassata et al 2011 (green squares). At redshift 3.0 we show the LF observed by Cassata et al 2015 (green squares) and Ouchi et al. 2008 (blue dots). At redshift 5.7 and 6.7 we show the LF observed by Ouchi et al. 2008 (blue dots) and Konno et al. 2018 (purple diamonds). }
\caption{ {\bf a)} The stellar mass - halo mass distribution at $z=3.0.$ The gray shaded region shows the distribution for the full \galform\sample. The solid yellow and blue lines and correspond to the median of\galform\central galaxies disk and bulges properties respectively. The shade regions show the 10-90 percentiles. The red dots show the{\it Thin Shell} LAE sample median, 10-90 percentiles (vertical) and the bin size (horizontal). {\bf b)} Same as a) but for the stellar mass - metallicity distribution. {\bf c)} Same as a) but for the stellar mass - star formation distribution. The top panels show the distributions of the halo mass, star formation and metallicity, respectively, for the full \galform\(yellow and blue for disk and bulge dominated respectively) and the {\it Thin Shell} model (red). The stellar mass distribution is shown in the right vertical panel.}
\caption{The \lya\\fesc\as a function of SFR (left panels) and stellar mass (right panels) at$z=3$. Gray points are from \citet{oyarzun17}. Each panel displays our model predictions with a different outflow geometry, as shown in the legend. The bottom-right corner displays the predictions of the model with no radiative transfer. The solid line in each panel is the median of \fesc\predicted by our models. The dark and light coloured shaded regions display the$32-68$ and $5-95$ percentiles of the models predictions, respectively.}
\caption{ {\bf Top}: the halo occupation distribution (HOD) at redshift 2.2 , 3.0 , 5.7 and 6.7 from left to right. Model with radiative transfer show as \colorWind , \colorBicone\and\colorThin\solid lines for the{\it Wind}, {\it Bicone} and {\it Thin Shell} geometry respectively. The LAE sample \NoRT\is plotted as dashed black line.{\bf Bottom}: fraction of galaxies that are considered LAE times the bias of the hosting dark matter halo. This quantifies the contribution of the different $\rm M_{h}$ to the overall bias of the population. }
\caption{ {\bf Top panels} : Monopole (3D auto-correlation function) for the \NoRT\sample (black), for the{\it Thin Shell} (\colorThin), galactic wind (\colorWind) and biconical galactic wind (\colorBicone) for redshift 2.2 , 3.0 , 5.7 and 6.7 from left to right. {\bf Middle panels} : The ratio between the different LAE sample and the dark matter correlation function. {\bf Bottom panels} : relative difference between the different samples and the \NoRT\monopole correlation function.}
\caption{ Distribution of the dust optical depth for the RT LAE samples for $z=2.2$, 3.0, 5.7 and 6.7 from left to right. Solid lines represent the {\it Thin Shell} (\colorThin), galactic {\it Wind} (\colorWind) and biconical galactic wind (\colorBicone) models. In black dash lines we show the $\tau_a$ value below which the typical discrepancies between our $f^{Ly\alpha}_{esc}$ model and the MC RT code are $<10\%$ ($\log \tau _a = -0.5$). In each panel we also indicate the percentage of LAEs with $\log \tau _a > -0.5$ }
\caption{GW amplitude $\sqrt{S_h}=2| \tilde{h} | \sqrt{f}$ of a black-hole binary source similar to GW150914 compared to the noise curves $\sqrt{S_n}$ of LISA \cite{2018arXiv180301944C}, LIGO \red{\cite{2016LRR....19....1A}, and a planned 3rd-generation detector \cite{2017CQGra..34d4001A} (both in their broadband configurations and with narrowband tunings). } Optimized {narrowbanding} % enhances (decreases) the detector sensitivity around the frequency $f_{33}$ ($f_{22}$) of the first subdominant (dominant) % mode of the BH ringdown. The BH binary waveform is generated using the approximant of \cite{2016PhRvD..93d4007K} with $m_1+m_2=65M_\odot$, $q=0.8$, $D=410$ Mpc, $\iota=150^\circ$ assuming optimal orientation ($\theta=\phi=\psi=0$).}
\caption{Top and middle panels show median values of $\delta {\rm GR}$ for LIGO at design sensitivity and with narrowband tuning, respectively; bottom panel shows the median gain $\zeta$. Data are shown as a function of total mass $m_1+m_2$ and mass ratio $q$ of the merging binaries; medians are computed over $\theta, \iota, \beta, \phi,$ and $\psi$. The distance is fixed to $D=100$ {\rm Mpc}. Binaries to the right of the dashed lines have sky-averaged LISA SNRs greater than 8 (these are computed following Ref.~\cite{2016PhRvL.116w1102S} using the updated noise curve of Ref.~\cite{2018arXiv180301944C} \red{and the nominal mission duration $T_{\rm obs}=4$ yr; the initial frequency is estimated such that the binary merges in $T_{\rm obs}$)}. Triangles indicate measured LIGO events \red{(we show the medians of the posterior distributions from \cite{2018arXiv181112907T})}.}
\caption{\redbf{$\textbf{(a)}$ RV oscillation amplitude and $\textbf{(b)}$ RV jitter due to stellar oscillations. In each panel, the bottom horizontal axis is \numax, while the top horizontal axis is the typical oscillation period (the reciprocal of \numax). The calculated values with different colors are separated with $\numax=500\ \muHz$, used for the subsequent model fitting. An over-density bump at \numax\$\sim$ 30 \muHz\arises from red clump stars.}}
\caption{\redbf{Comparison of the \textit{calculated} with the \textit{predicted} RV jitter \sigmarv\(see the text for their definitions) using three models as indicated. Grey dashed lines represent perfect agreement. We separately fitted both dwarfs and subgiants (blue diamonds), and giants (red squares) using a dividing point $\numax = 500\ \muHz$, or equivalently \logg\$\sim$3.5 dex. The fractional residuals are defined as ($\rm \sigma_{rms, RV, calc}-\sigma_{rms, RV, pred})/\sigma_{rms, RV, pred}$. The bump at \sigmarv\$\simeq$ 4 m/s\is caused by red clump stars. Black squares indicate stars with long RV time series from which we computed\sigmarv\and rescaled it to include contributions only from oscillations (see the text). Here we do not show the comparisons for\amp, \vosc, and \sigmaphot\given their almost identical features.}}
\caption{\redbf{Stellar parameters for 21 bright stars with RV time series observed by ground-based telescopes.}}
\caption{$\rm log$ $g$ vs. \teff\diagram color-coded by the RV jitter\sigmarv. Approximate \numax\is labeled in the right vertical axis. The solid lines show evolutionary tracks from PARSEC\citep{bressan12}, with the masses from 0.8 to 2.0 $\rm M_{\odot}$ and the metallicy [Fe/H]\=\-0.096 equal to the median value of the whole sample.}
\caption{Time evolution of the dimensionless kinetic energy for $Wi = 1$, $\beta = 0.5$ and $Re=0.01$. The simulations have been performed with two uniform grids; M1 and M2 (64$\times$64 and 128$\times$128, respectively) and an adapted grid. Numerical results with (i) the log-conformation kernel and (ii) the square root kernel are shown. Results of Fattal \& Kupferman (2005) are also shown (red circles). Insert: Sketch of the lid cavity problem.}
\caption{Diagram of GRAD architecture. {\color{red} Red} connection indicates normal forward propagation, but back-propagation will reverse the signs. % }
\caption{Difference of transmittance \(\Delta T\) between the isotropic and anisotropic thin film of graphene for a system air/graphene/glass. The refractive index of glass is taken as \(1.5\). Graphene is modeled using the Kubo formula with \(E_F=0.4\)\,eV,\(\tau=100\)\,fs. The red dotted line represent a thickness equal to\(1/1000\) of the wavelength.}
\caption{% Values of \(\mathrm{Re}[\chi^s_x]\) for \MoS2 single layer retrieved with the isotropic thin film model (\textcolor{blue}{x}) and the anisotropic current sheet model (\textcolor{black}{\(\square\)}). Data from~\cite{Jayaswal_18}. The (\textcolor{red}{\(\diamond\)}) curve is calculated from (\textcolor{blue}{\(\times\)}) using the shift calculated from (\ref{Eq:chia_from_chiI}) (\textcolor{green}{\(+\)} in the inset); (\textcolor{magenta}{\(\Delta\)}) is (\textcolor{blue}{x}) shifted by \(-\varepsilon^b_x d\) (\textcolor{cyan}{\(o\)} in the inset). The dielectric function of N-BK7 glass (substrate used in~\cite{Jayaswal_18}) is taken from Sellmeier's equation provided by Schott~\cite{rii-BK7}.}
\caption{ Ratio of the fitting variables of the basis function on the FS ``$f$", from the \blue{$d$-wave-like gap function.} We take the \blue{intra-unit-cell} component \blue{$C_0^f$} as unity; to stress the different sign between the two bands, we put different signs. }
\caption{ Ratio of the \blue{component of the well-known SC gap on the both FSs based on Table~\ref{table2}.} }
\caption{Mean Absolute Error of different methods on the WorldExpo’10 dataset. Our method achieves superior performance with respect to the average MAE of five scenes. The best results and the second best results are highlighted in {\textcolor{red}{red}} and {\textcolor{blue}{blue}}, respectively. Best viewed in color.}
\caption{The structure of the dataset where the \textcolor{red}{red text} is verbatim.}
\caption{The organization of the pouring data where the \textcolor{red}{red text} is verbatim}
\caption{\color{Gray} \textbf{Spike-rate definition.}. There are several definitions of spike rate. Averaging over all the spikes emitted by a single neuron (spikes in the red box), we get the \emph{spike-count rate}. Averaging over the spikes emitted at a time instant by all the neurons (spikes in the green box), we get the \emph{instantaneous population rate.}}
\caption{\color{Gray} \textbf{N-MNIST time collapsed images.}. N-MNIST patterns are collapsed in the time dimension to static images with pixel intensity proportional to the event rate of the pixel. These images are trained on an ANN to examine the removal of the temporal component in N-MNIST affects the performance. The above are 6 such images created from N-MNIST time-collapsed patterns.}
\caption{\color{Gray} \textbf{Spiking Neural Network Architecture.}. (Modified version of Figure 2 in \cite{Iyer:2017}) The N-MNIST patterns, each pattern representing a \emph{saccade} project to the excitatory layer. There is an all-to-all connection from the input to excitatory layer. The excitatory layer has a one-to-one connection with the inhibitory layer. Upon firing, an excitatory neuron activates its corresponding inhibitory neuron, which in turn inhibits all excitatory neurons except the one it received excitatory connections from. Top right: Kernel functions for the currents in the network - $I_{xe}$ has a gradual rise in current followed by a gradual fall. The other two currents, $I_{ei}$ and $I_{ie}$ have an instantaneous rise in current followed by a gradual fall.}
\caption{\color{Gray} \textbf{Two regimes of STDP operation.} By varying the decay time constant of the presynaptic spike trace $\tau_{xpre}$ we get two regimes of STDP operation, (A) When $\tau_{xpre}$ is low, the presynaptic trace ($x_{pre}$) decays quickly. The value $x_{pre}$ at the time of the postsynaptic spike (green dot) depends on the time difference between the pre- and post-synaptic spikes. In this regime, spikes that occurred much earlier than the postsynaptic spike time have no impact on learning. (B) When $\tau_{xpre}$ is high, $x_{pre}$ decays slowly. At the time of postsynaptic spike, $x_{pre}$ (green dot) depends on the number of spikes (i.e. spike-count rate) alone. Precise presynaptic spike times do not have much impact on learning. }
\caption{\color{Gray} \textbf{Results on \emph{log scale}.}Design Space Exploration results, all parameters are plotted on a $log$ scale, trained over one epoch. Results progressively get better as $\tau_{xpre}$ increases. The best overall result (82.46\%) is marked by a red star in the top right. The best overall results, and a larger number of better results ($>80$\%) occur at higher values of $\tau_{xpre}$. The best result for the low $\tau_{xpre}$ (= 20 ms) is marked by a red triangle on the top right. }
\caption{\color{Gray} \textbf{Results on \emph{linear scale}.} The best result in the \emph{log scale}, (Figure \ref{fig:results}, $\tau_{xpre} = 215 ms, \eta = 0.05, \Delta \theta = 0.1 mV$) was taken. $\tau_{xpre}$ was fixed at $215$ ms and values of the other parameters ($\Delta \theta$ and $\mu$) around the best result were plotted on linear scale to determine if there were better results around the parameter space. Training was done over one epoch. We see that generally results get better as learning rate ($\eta$) gets higher and spike frequency adaptation rate ($\Delta \theta$) gets higher. The best results are marked with a red star on the top right. We get a slight improvement in accuracy, $84.51\%$. }
\caption{\color{Gray} \textbf{Weight distribution} plotted after training for one epoch (left) and training for 4 epochs (right). This is plotted for the best parameter values for $\tau_{xpre} = 20$ ms (Threshold adaptation rate, $\Delta \Theta = 0.001 mV$, Learning Rate, $\mu = 0.005$) and $\tau_{xpre} = 200$ ms (Threshold adaptation rate, $\Delta \Theta = 0.01 mV,$ Learning Rate, $\mu = 0.05$). After running more epochs, the networks become more stable. }
\caption{\color{Gray} \textbf{Testing accuracy over epochs.} Testing accuracy is plotted as a function of number of training epochs. This is plotted for the best parameter values for $\tau_{xpre} = 215$ ms (blue) (Threshold adaptation rate, $\Delta \Theta = 0.01 mV,$ Learning Rate, $\mu = 0.05$) and $\tau_{xpre} = 20$ ms (red) (Threshold adaptation rate, $\Delta \Theta = 0.001 mV$, Learning Rate, $\mu = 0.005$). As can be seen, results improve over time when $\tau_{xpre}$ is high, but actually decrease over time when $\tau_{xpre}$ is low. }
\caption{\color{Gray} \textbf{Weights over epochs.} Weights obtained after training on one epoch (left) and 4 epochs (right) for $\tau_{xpre} = 20$ ms (top row) and $\tau_{xpre} = 215$ ms (bottom row) are plotted. The weight for each neuron is plotted on a 34 $\times$ 34 grid. For the low $\tau_{xpre}$ (= 20 ms) case, while the digits can be vaguely discerned after training on one epoch, after training for 4 epochs, only part of the digit is visible. (For e.g. Digit ``2'' on the top right corner and digit ``5'' on the bottom left corner.) On the other hand, for the high $\tau_{xpre}$ case (= 215 ms) digits become slightly more discriminative when training on more epochs. (For e.g. digit ``3'' third row far right. }
\caption{\color{Gray} \textbf{STDP function derived from the training data.} \emph{Left:} $H(t)$ is the average number of presynaptic neurons spiking at instantaenous time $t$ in a pattern. The function $H(t)$ derived from all the training patterns is given in the figure. \emph{Right:} $h(t)$ is the STDP curve obtained after scaling and biasing $H(t)$ to ensure stability and to preserve the learning dynamics described in the previous sections. }
\caption{\small (a) Epifluorescence image of a flk1:GFP zebrafish embryo, showing the region of interest (ROI) used in our experiment. CV: cardinal vein, CA: cardinal artery, ISV: intersegmental vessel. (b) PSF encoded image showing 1 µm fluorescent tracer beads flowing within the ROI. (c) 3D trajectories of tracers within cardinal and intersegmental vessels of zebrafish, reconstructed from 2000 frames. Color-coding is used to distinguish the trajectories of each tracer particle. Supplementary \textcolor{blue}{Visualization 1} shows the tracers flowing with the blood in real-time, along with a 3D reconstruction of tracer trajectories.}
\caption{\small Composite images showing detected tracer locations (green spheres) embedded within a surface rendering (red) of the 3D blood vessel structure obtained from quasi-simultaneous SPIM imaging (Dataset 1\cite{ZhouData2018}). (a) View of several tail segments; (b) detail of a vessel junction. The recovered bead coordinates lie within the blood vessels, demonstrating that the Airy-CKM method is correctly identifying the 3D coordinates of the beads. Supplementary \textcolor{blue}{Visualizations 2 \& 3} show multi-view animations of the composite image and beads flowing within SPIM-reconstructed blood vessels.}
\caption{Results on SESIV. The best results are shown in \textcolor[rgb]{0,0,1}{blue}.}
\caption{Effectiveness of confident instance utilization. The best results are shown in \textcolor[rgb]{0,0,1}{blue}.}
\caption{Effectiveness of identity tracking. The best results are shown in \textcolor[rgb]{0,0,1}{blue}.}
\caption{Some examples of calculated priority values. In this example the weight factors are $a=1$, $p=10$, and $t=10$. The final priority value is constructed using Equation~\ref{eq:priority}. The elements are coloured matching which part of the priority value they form - the rank (\textcolor{red}{red}) and repeat number (\textcolor{orange}{orange}) form the integer part of the priority, the ToO flag (\textcolor{violet}{purple}) forms the first decimal place and the tie-break parameters (airmass, probability and time-since-observed, \textcolor{teal}{teal}) combined form the later decimal places. }
\caption{ \textcolor{red}{Todo: Draw the curve.} Precision at $k$ (P@$k$, \%) of different methods evaluated at top $k$ returned images on the CompCars, eBayCamera10k, and Lookbook dataset.}
\caption{\textcolor{red}{[Draw a new one]} The architecture of the proposed adversarial network for learning discriminative features for fine-grained image retrieval. Our network consists of a generator $G$ and an evaluator $E$. While $E$ aims at distinguish the generated representations from the real images, we train $G$ to generate image representations that can confuse $E$. $G$ and $E$ are trained alternatively in an adversarial manner until convergence.}
\caption{\label{tb:internal-gw} Full length \textsc{Rouge} F1 evaluation on English Gigaword test set of \citet{zhou-EtAl:2017:Long} and our internal test set. Results with \otherpaper{} mark are taken from the corresponding papers. The ABS baseline fails to run with the latest stable \red{Torch} framework, so we omit its performance on our internal test set. }
\caption{Rarefaction test case for expansion fan. The graph on the right shows profiles of axial velocity and density for the Riemann problem in Figure~\ref{riemann} at $t=0.01$, the initial condition chosen for the rarefaction flow. {\color{red}The vertical dash line at $x=0.5$ indicates} the position of the diaphragm locating the discontinuity in Figure~\ref{riemann}. The outflow boundary coincides with the diaphragm; the actual shock and contact discontinuities are not present in the domain. \textcolor{red}{The variation in particle distribution is due to the inherent random distribution and number of particles. Since particles are not effecting the fluid or each other in this test, particles were allowed to overlap with one another which leads to higher than expected variation in the distribution here.}}
\caption{Left and right states (relative to \textcolor{red}{the vertical dash line at $x=0.5$} in Figure~\ref{fig:expfan}) for the Riemann problem used to initialize the flow.}
\caption{Block entanglement entropy for $J=0.9U$ and $L=226$ as a function of the chord length $\ln(d(m|M))=\ln\left(\left[\frac{M}{\pi}\sin\left(\frac{m\pi}{M}\right)\right]\right)$ and compared to the CFT prediction Eq.~\eqref{blockentropy} for \textbf{(c)} the non-frustrated case ($\phi=0$) and \textbf{(a)} the fully frustrated case ($\phi=\pi$). \textbf{(b)} is an illustration of the different cuts that can be made on the model. The cuts after A,B and C are distinguished using the symbols $\ast,+, \star$ respectively. } \label{entropy} \end{figure} %%%%%%%%%%%% For a critical system with open boundary conditions, conformal field theory (CFT) predicts that the von Neumann entanglement entropy scales as: \begin{equation}\label{blockentropy} S_L(\ell)=\frac{c}{6}\ln\left[\frac{L}{\pi}\sin\left(\frac{\pi \ell}{L}\right)\right] + A + \mathcal{O} \left(\frac{1}{\ell}\right) , \end{equation} where $c$ is the central charge, which can be used as an indicator of the universality class of the corresponding field theory, and $A$ is a model dependent (i.e., non-universal) constant~\cite{Vidal2003,Calabrese2004}. Here we calculate this based on which cell each site occupies, so $L$ is replaced by $M$ in Eq.~\eqref{blockentropy} and $m$ is used to show which cell we are cutting. In Fig.~\ref{entropy} we distinguish three different cuts of the chain, according to the sub-lattice after which they take place (see \ref{fig:modelcut}), and perform the fit of Eq.~\eqref{blockentropy} on each separately. Data are shown for $J/U=0.9$ and we introduced the chord distance $d(m |M)=\left[\frac{M}{\pi}\sin\left(\frac{m\pi}{M}\right)\right]$ for convenience. The $C$-cut splits a rhombus in half and therefore gives rise to a higher entropy with respect to the $A,B$-cuts. Alternatively, we can understand this by considering that the $C$-cut separates two cells and that correlations have a strong oscillatory character between neighbouring cells, causing a supplementary amount of entanglement. For the LL of the unfrustrated regime $\phi=0$, in Fig.~\ref{entnonfrust}, we find that the cut after the maximal cut C has the central charge $c_C^{[\phi=0]}= 1.0180 \pm 0.0003$. The cuts after $A$ and $B$ have central charges $c_A^{[\phi=0]}=1.101 \pm 0.001$ and $c_B^{[\phi=0]}= 1.050 \pm 0.001$ respectively. For the PLL of the frustrated regime $\phi=\pi$, in Fig.~\ref{entfrust}, we find that the maximal cut fits such that the central charge $c_C^{[\phi=\pi]}= 1.052 \pm 0.002 $, which is comparable to the unfrustrated case. For the cuts after $A$ and $B$ the central charges are $c_A^{[\phi=\pi]}=1.142 \pm 0.006$ and $c_B^{[\phi=\pi]}=1.078 \pm 0.005$ respectively. %%%%%%%%%%%%%%%%%%%%% From following section. \begin{figure*}[t!] %\begin{subfigure}{0.495\textwidth} %\includegraphics[width=6.5cm,trim={2.2cm 0.5cm 0.6cm 0cm}]{fig11a.png} %\caption{} %\label{ESCnumbermixed} %\end{subfigure} %\begin{subfigure}{0.495\textwidth} %\includegraphics[width=5.5cm,trim={1.5cm 0.5cm 2cm 0.31cm}]{fig11b.png} %\caption{} %\label{ESunfrustTL} %\end{subfigure} %%\vspace{3cm} %\begin{subfigure}{0.7\textwidth} %\includegraphics[width=0.95\columnwidth,trim={2.2cm 0.1cm 0.6cm 0cm}]{ESCnumbercurve.png} %\caption{} %\label{ESCcurve} %\end{subfigure} \includegraphics[width=0.78\textwidth,trim={0cm 0cm 0cm 0cm}]{fig11.png} \caption{\textbf{(a):} The entanglement spectrum as a function of the dispersion from uniform filling $\delta N$ of the number of bosons for the cut after $C$. $\phi=0$ denotes the unfrustrated case ($\circ$), $\phi=0.9\pi$ the intermediate frustration ($\square$) and $\phi=\pi$ denotes the fully frustrated case ($\star$), simulated at $J=0.9U$. In \textbf{(a)} a solid line is used to join the degenerate eigenvalues in all cases. \textbf{(b):} Approximate parabolas for $\phi=0$, $\phi=0.9\pi$ and $\phi=\pi$ (left to right) based on the length $L=226$. In the third panel different colours denote the possible curve fitting to even and odd. \textbf{(c):} The unfrustrated ES for a cut after $C$ at $L=226$ with the thermodynamic limit approximation shown by the parabolas. The legend shows the degeneracy of each parabola.} \label{ESplot} \end{figure*} %%%%%%%%%%%%%%%%%%%%% Values for both the frustrated and the unfrustrated are thus fully compatible with the well-known result for the LL phase of the Bose-Hubbard model on a purely 1D chain~\cite{Vidal2003, Calabrese2004, Ejima2012}, i.e., $c=1$. This confirms that only one bosonic component (out of three possible ones) becomes gapless, in either case, as we have already seen via the correlations in the previous section. This holds regardless of which cut in the system is fitted, i.e. even if we fit every cut after A or B which is a cut across cells we get $c\simeq 1$, once we have considered that finite size effects are taking place. We, therefore conclude that $c$ is not a good indicator to distinguish PLL from LL. The low-lying levels of the entanglement spectrum, however, may allow this as shown in the next section.\\ Before turning to the entanglement spectrum analysis, let us mention that the entropy scaling across the PLL-LL transition at finite deviations from $\phi=\pi$ is not displaying any clear signature of a $c=3/2$ CFT line, as one would expect from its predicted Ising character~\cite{Doucot2002,Ioffe2002}. The difficulties in analysing transitions between gapless phases has already been noticed in spin models\cite{DeChiaraPRB2011}. \subsection{Entanglement Spectrum} Despite having the same central charge, we expect qualitative differences between the wavefunction structure inside the LL and PLL phase. We, therefore, resort to the entanglement spectrum, which is capable of revealing key properties about the system, such as symmetries and excitations, which the von Neumann entropy, being a single number, is unable to provide ~\cite{LiHaldane, Oshikawa, ESDeng, DeChiaraPRL2012,ESDeChiara2013}. % Here we choose to focus on the $C$-cut which leaves $(M-1)/2$ rhombi on each side, so that the bipartition is perfectly symmetric, at least concerning the number of sites. Thanks to the conservation of the total number of particles in the system, each Schmidt eigenvalue $\lambda_i$ can be associated to an eigenvector of the reduced density matrix with a fixed number of particles. In Fig.~\ref{ESplot}\blue{a} we plot the $\lambda_i$'s for a chain of $75$ rhombi, according to the excess number of particles $\delta N$ with respect to a homogeneously distributed unit filling (i.e., $L/2$ particles on both sides of the bipartition), similarly to Ref.~\cite{Lauchli2013Operator}. The tunnelling value we consider is $J/U=0.9$, inside both the LL and PLL phases (see Fig.~\ref{phasefill1full}). The eigenvalues are clearly symmetric with respect to $\delta N=0$ regardless of the amount of frustration. For the unfrustrated LL at $\phi=0$ (blue $\circ$) and the intermediate frustrated LL at $\phi=0.9\pi$ (green $\square$), it is easy to recognise both the $-\log\lambda_i \propto \delta N^2$ dependence and also the starting of the equally spaced CFT tower within each distinct $\delta N$, as predicted for the standard 1D Bose-Hubbard chain~\cite{Lauchli2013Operator}. Both features apparently disappear for the PLL at full frustration $\phi=\pi$ (red $\star$), thus signalling a dramatic change in the underlying wavefunction, undetected by the entropy scaling analysis. In order to examine this more clearly we plot fitted curves at length $L=226$ of the same curvature for given $\phi$ in Fig. \ref{ESplot}\blue{b} for $\phi=0$, $\phi=0.9\pi$ and $\phi=\pi$ from left to right. It is evident from this that the unfrustrated cases can be extrapolated to the typical curves. For $\phi=\pi$, however, it is impossible to fit the eigenvalues with functions of the same curvature. For example, if the first 5 points are examined closely it can be seen that a parabola would not be able to fit adequately both 1 to 2 and 3 and 1 to 4 and 5. Instead it seems that two distinct parabola sets are appearing at the even and odd $\delta_N$'s as shown by the red and black curves. In Fig.~\ref{ESplot}\blue{c} we present the results of a finite-size scaling towards the thermodynamic limit for the unfrustrated case ($\phi=0$) shown by the parabolas. A modified degeneracy counting and the appearance of a secondary tower, both possibly related to the internal structure of the lattice, are evident. Examining these at $\delta_N=0$ the spacing of the parabolas between every second one is approximately equal i.e., $1-3 \approx 2-4 \approx 3-5$, whereas the spacing between neighbouring parabolas differs. Higher parabolas are excluded as they fall below the accuracy of our results. For the fully frustrated case, instead, where it seems that two distinct parabola sets are appearing, a reasonable finite-size scaling procedure is not possible without conserving explicitly the local $\mathbb{Z}_2$ quantities. Both these aspects go beyond the scopes of the present work and deserve future investigation. We do, however, notice that the resulting pattern of (quasi-) degenerate multiplets in the entanglement spectrum changes quite radically in the fully frustrated case from the unfrustrated and intermediate frustrated cases. So we are confident that the entanglement spectrum can be used to distinguish between the two cases, where the entropy cannot. %%%%%%%%%%%%%%%%%%%%%% \section{Conclusions} \label{sectconc} In this paper we have analysed the ground state phase diagram of a system of interacting bosons on a geometrically frustrated lattice; namely, we have considered a quasi-1D chain of rhombi pierced by magnetic flux. % For unit filling and a sufficiently low tunnelling amplitude the system is in the Mott insulator phase as expected. For larger tunnelling values, we have numerically confirmed that when full geometric frustration prevents the movement of single particles (i.e., the bands become flat), the system still enters into a gapless phase where the elementary moving objects are pairs of particles. We have explored the regime where the frustration is not perfect, highlighting that the pair fluid can only be obtained for a very small region, making this quite challenging for an experimental realisation, especially at low particle filling. It is, however, possible to extend this region by a small amount using higher filling and even more by using amplitude modulation within the system instead of the phase shifts we applied, which is also possible experimentally. % From a different perspective, we have highlighted that, whilst the central charge obtained from the entropy cannot be used to distinguish between the PLL and LL phase, the features and quantum numbers of the entanglement spectrum do have noticeable differences between the two. There are a number of directions this work could be expanded upon, for example: i) compare the robustness of the PLL phase with respect to other deformations of the flat bands (such as amplitude modulation), and compare it to other flat band models (e.g., the Creutz ladder~\cite{MRizzi2017,Tovmasyan2018}), to see whether the (here hidden~\cite{Kremer2018}) topological character plays any role; ii) work out an explicit mapping to the effective Ising model predicted by Dou\c{c}ot and Vidal~\cite{Doucot2002}, in terms of measurable quantities (as done for Creutz ladder fermions~\cite{MRizzi2017}), in order to shed new light on the nature of the PLL-LL transition (possibly once the PLL region is also extended to simplify things); iii) deepen the understanding of the striking change in the entanglement spectrum, possibly by also explicitly enforcing the extensive number of $\mathbb{Z}_2$ symmetries~\cite{Doucot2002,Tovmasyan2018} in the numerics.\\ Moreover, it would be very interesting to examine the dynamics of our interacting chain, in order to formulate experimental detection strategies, now that platforms for artificial flat-band systems are flourishing again~\cite{Leykam2018,Kremer2018,Mukherjee2018}. \acknowledgments The authors would like to acknowledge useful discussions with K. McAlpine, A. Trombettoni, F. Gerbier, B. Dou\c{c}ot, J. Vidal, S. Al-Assam and A. Haller. % The authors are grateful for the computational time from the Mogon cluster of the JGU (made available by the CSM and AHRP), and S. Montangero for a long-standing collaboration on the flexible Abelian Symmetric Tensor Networks Library employed here. % CC wishes to thank the EPSRC and the Professor Caldwell Travel Studentship for support.\\ \bibliographystyle{apsrev4-1} \bibliography{ref} \vspace{0.5cm} \appendix* \renewcommand\thefigure{A.\arabic{figure}} \section{Increasing the robustness of the PLL phase} \manuallabel{sec:app}{Appendix} \setcounter{figure}{0} In order to increase the robustness of the PLL phase the Hamiltonian can be formulated using a tunnelling modulation instead of the magnetic flux used previously (see Sec.~\ref{sectmod}). In order to do this the Hamiltonian is again given by: \begin{eqnarray} \label{BHeqncos} \hat{H}_{BH}&=& \hat{H}_0 + \hat{H}_U \equiv \hat{H}_0 + \frac{U}{2}\sum_j \sum_{\alpha} \hat{n}_{j,\alpha}(\hat{n}_{j,\alpha}-1) \\ % \hat{H}_0 & = & - J \sum_j \sum_\ell \sum_{\alpha,\beta} T^{(\ell)}_{\alpha ,\beta} \\hat{b}^{\dagger}_{j+\ell,\alpha} \hat{b}^{\phantom{\dagger}}_{j,\beta} \,\end{eqnarray}\\ %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[h!] \includegraphics[width=0.97\columnwidth,trim={0.05cm 1.1cm 0.05cm 2.5cm}]{app-fig1.png} \caption{The phase diagram for an infinite estimated length of the Hamiltonian Eq.~\eqref{BHeqncos} for amplitude modulation of $\cos(\phi)$ against parameters $J/U$. The MI, LL and PLL regions are as labelled.The MI-LL and the MI-PLL transitions are again obtained from the compressibility of the energy gap (see Sec.~\ref{sectphase}). The LL and PLL phases are characterised by the decay of the correlation functions as illustrated in Sec.~\ref{sectPLL}).} \label{appcosphase} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% where the labels are as previously defined (Sec.~\ref{sectmod}). %where the index $j$ denotes the lattice cell, the Greek letters label the basis inside a cell, i.e., $\alpha, \beta \in \{A,B,C\}$, and $\ell \in \{0, \pm 1\}$ represents the (relative coordinate of the) cell where the particle is hopping to. %We use the annihilation (creation) operators $\hat{b}^{(\dagger)}_{j,\alpha}$, %and the number operator $\hat{n}^{\phantom{\dagger}}_{j,\alpha}=\hat{b}_{j,\alpha}^{\dagger} \hat{b}^{\phantom{\dagger}}_{j,\alpha}$. In order to modulate the amplitude, a $\cos(\phi)$ factor is included in the $C-B$ leg of each diamond (as shown by the dashed lines in Fig.~\ref{model}). This means the hopping matrices are now instead: \begin{equation}\label{eq:tunnellcos}{T}^{(0)} = \left(\begin{array}{ccc} 0&1&0\\ 1&0&1\\ 0&1&0\\ \end{equation}\end{array} \right), \quad{T}^{(+1)}=\left(\begin{array}{ccc} 0&1&0\\ 0&0& \cos(\phi)\\ 0&0&0\\ \end{array} \right), \quad{T}^{(-1)}= \left({T}
\caption{Active Road Segments. Road segments colored in {\color{red}red} are active}
\caption{Multifrequency light curve of CTA 102 obtained for the period from 2008 August to 2018 January. {\it a)} \gray{} light curves with adaptive (red; $\geq156.1$ MeV) and 2-day (blue; $100$ MeV) bins, {\it b) and c)} the flux and photon index with 2- and 7-days binning, {\it d)} Swift XRT light curve in the 0.3-10 keV range, {\it e)} UV/optical fluxes in $V$, $B$, $U$, $W1$, $M2$ and $W2$ bands and {\it f)} the energy and arrival times of the highest-energy photons. The vertical blue dashed line shows the period when a large flare in the $R-$ band was observed (28 December 2016). }
\caption{CTA 102 \gray{} photon index vs. flux in adaptive (orange) and two-day bins (blue). Similar plot for the X-ray band is shown in the insert.}
\caption{Density number of galaxies (black, left axis) and their cumulative sum (dashed-blue, rights axis) generated by \colore corresponding to 10 years of LSST data-taking.}
\caption{Summarisation results (\%) on TVSum and CoSum. $1^{st}/2^{nd}$ best in \textcolor{red}{red}/\textcolor{blue}{blue}. Full model means $r^g + r^l + r^u$.}
\caption{Modifications of different atomic line intensity due to GSA, the influence of anisotropic pumping and magnetic realignment. Line colors represent: {\color{blue} \bf Blue} for [C\,{\sc i}]$\lambda610{\mu}m$, {\color{RedOrange} \bf Orange} for [C\,{\sc ii}]$\lambda157{\mu}m$, and {\color{YellowOrange} \bf Yellow} for ${\rm C^{18}O(1-0)}$. The raw observational data are in dashed lines from \citet{Kamegai03}. The solid lines are the modifications purely due to anisotropic radiative pumping. The error bars that depict the possible range of the original column density of the corresponding element are produced accounting for the magnetic realignment. Potential peaks for neutral Carbon for both stripes are noted with green dotted lines. The intensity peaks for atomic and single-ionized Carbon lines are marked on the corresponding $y-$axis. }
\caption{Redshift distribution of the background sources that are lensed by the galaxy cluster \cname\are plotted. The gray area represents the training set (data available to the lensing groups when computing the models in 2015 from\citealt{Jauzac:14,Richard:14,Treu:15}) and the white is the new set (data available from the resent follow-up survey by \citet{Caminha:17} that did not have any previous archival spectroscopic redshifts). From the distribution, it is clear that the new arcs tend to have a higher redshift than those arcs in the training set. This color convention is going to be carried through the rest of of the paper.}
\caption{\hst\optical F160W image of\cname\where the colored marks represent the position of the multiply-images, and the color indicates the background-source-measured spectroscopic redshift\citep{Jauzac:14,Richard:14,Treu:15,Caminha:17}. The plus symbols represent the training set while the diamonds indicate the new lensed systems. The boxes indicate lensed system 26, which had a spectroscopic redshift update from $z_{26} = 2.1851$ to $z_{26} = 3.2355$.}
\caption{\hst\optical F160W image of\cname\where the positions of the multiple images are indicated by the circles, while the color represents the source plane scatter of the lensed systems. We find models were the RMS scatter is uniform across the image plane, while some others have regions of relative high and low RMS scatter. The square symbols represent the location of the lensed system 26, and the square or circle with a line across the center, e.g.,$\oslash$, represent ``failed'' arcs (see Section~\ref{sec:ippa}).}
\caption{Source plane RMS scatter plotted for each of the lensed systems using the version 4 publicly available lens models. All of the constraints are considered as training, since they were used as inputs when computing the new lens models. The blue ``x'' represents the mean of the distribution. In the text of each panel you can find the name of the algorithm used; in parenthesis if the algorithm is parametric (Par), non-parametric (NonPar), or hybrid (Hyb); the new constraints used, and additional comments to describe the difference between models from the same lensing group (two sets of galaxy cluster member catalogs, Cat1: \citealt{Grillo:15} and Cat2: \textcolor{blue}{Richard et al., in prep.}). The '*' in the name is to highlight the new teams that have made V4 lens models that did not compute a V3 lens models for \cname.}
\caption{\textbf{Consecutive foot steps and CoM trajectories in Cartesian and phase space} : \mycom ~and \mybluecircle ~are a current CoM and foot step, followed by red color representations. For each stances, a CoM \myapex ~above sagittal foot placement is called apex and the intersection \myswitch ~is represented by switching. $t_{switch}$ and $t_{apex}$ are defined as time duration from \mycom ~to \myswitch ~and from \myswitch ~to \myapex.}
\caption{Magnitude of TT\,Hor (green dots) in V filter after pre-whitening compared to the{\sc phoebe} model (black line). The residuals from the model are shown below.}
\caption{\red{HR diagram showing the evolutionary tracks of \textsc{mesa} models for the accretor and donor with different initial conditions as given in Table~\ref{tab:MESA_models}. Tracks for single stars (SS) of the same final mass as the accretor for each model are also shown. The inverted triangles indicate the current evolutionary state of each star. Error bars are shown for the accretor and donor from the \textsc{phoebe} binary model. The black stars indicate start of rapid mass transfer. The open circles indicate the equivalent evolutionary conditions in the $T_{\rm eff}$/$\log $($L$/L$_{\sun}$) parameter space for the single stars. }}
\caption{\red{Echelle diagram comparing scaled, calculated frequencies for Model B (upper panel) and the unscaled frequencies for Model D (lower panel). A large separation of frequency modulo\,=\,5.2 d$^{-1}$ is used in both cases. The matching observed frequencies for TT Hor are indicated by the coloured stars with brown = \textit{f}$_1$; green = \textit{f}$_2$; blue =\textit{f}$_3$; red = \textit{f}$_4$. For each plot, The highest radial order shown is \textit{n} = 10.} Positive $m$ values are prograde modes travelling in the direction of rotation.}
\caption{Identification of the modes of the observed frequencies in the accretor by comparison to modes identified by \textsc{gyre} \red{after applying a scaling factor of either 2.7\% (Model B) or no scaling factor (Model D).} }
\caption{\label{fig:grafana_multiblade} Grafana dashboard used \red{for monitoring packet and event related counters for} an implementation of Multi-Blade data processing.}
\caption{(a) Cumulative Error Distribution curves and (b) qualitative results obtained by the {\color{blue} \textbf{CSTN-5}} and {\color{green} \textbf{DeSTNet-5}} on IDocDB.}
\caption{Qualitative results obtained with {\color{blue} \textbf{CSTN-5}} and {\color{green} \textbf{DeSTNet-5}} on IDocDB. (Results are best viewed on a digital screen)}
\caption{Qualitative results obtained with {\color{blue} \textbf{CSTN-5}} and {\color{green} \textbf{DeSTNet-5}} on IDocDB. (Results are best viewed on a digital screen)}
\caption{ The training log of a 7,500-landmark classifier. % The boosting algorithm successfully reduced the training cost (\textcolor{blue}{blue}) at each iteration. % The precision@1 (\textcolor{red}{red}) of the evaluation set classification steadily grows to reach the level of 72.8\%. % The feature vectors contain 16,000 context dimensions (1,000 regions with 16 visual words each) and 384 descriptor dimensions. % Overall, 985 context features and 1,015 descriptor features were selected by the algorithm. % Some of the feature dimensions were used by multiple decision stumps and can be combined to speed up the classification. }
\caption{ Comparisons of our proposed approach to the state-of-the-art on PRID2011, iLIDS-VID and MARS datasets. The rank1 accuracies are reported and for MARS we provide mAP in brackets. The best and second best results are marked by {\color{red}{red}} and {\color{blue}{blue}} colors, respectively. }
\caption{Test accuracies on \mini\and\tiered. For each dataset, the first set of results use convolutional networks, while the second use much deeper residual networks, predominantly in conjuction with pre-training. }
\caption{Architecture details for 5-way 1-shot \mini\and\tiered. The shapes correspond to the meta-training phase. We used a meta-batch of $12$ task instances in parallel.}
\caption{Learning rate annealing schedules used to train feature extractors for \mini\and\tiered.}
\caption{Architecture details for 5-way 1-shot \mini\and\tiered. The shapes correspond to the meta-training phase. We used a meta-batch of $12$ task instances in parallel.}
\caption{Model parameters adopted in the analysis: fiducial values and priors used throughout unless otherwise stated. Values for the cosmology sector and X-ray luminosity--mass and temperature--mass relations are taken, respectively, from \textcolor{blue}{Komatsu et al. (2009)} and \textcolor{blue}{Vikhlinin et al. (2009a)}. $\mathcal N ( \sigma ) $ represents the normal distribution with mean given by the fiducial value and variance $\sigma^2$; $\mathcal U(x_1,x_2)$ denotes the uniform distribution with endpoints $x_1$ and $x_2$; and $\mathcal \delta_D[x] $ signifies that the parameter is kept fixed at the fiducial value $x$. Throughout the paper, we assume a flat cosmology, with $\OL = 1 - \OM$; Gaussian initial conditions, $\FNL \equiv 0$; and temperature--mass relation parameters by \textcolor{blue}{Vikhlinin et al. (2009)}, coupled to the luminosity--mass relation via a bivariate lognormal distribution with null correlation coefficient between luminosity and temperature (see \textcolor{blue}{Pillepich et al. (2012)} for a discussion). Standard priors are taken from \textcolor{blue}{Riess et al. (2011)} for the Hubble Parameter ($\Delta h = \pm 0.022$), \textcolor{blue}{Cooke et al. (2014a)} for the mean baryon density from primordial deuterium measurements coupled to standard models of Big Bang nucleosynthesis ($\Delta (\OB h^2) = \pm0.0045$), and from \textcolor{blue}{Vikhlinin et al. (2009a)} for the luminosity--mass parameters.}
\caption{\label{TAB:ERO_ANALYSIS}Scenarios for the analysis of the $\eRO$ data. `Self calibration' denotes the configuration adopted for the main results of {\color{blue}{Pillepich \etal 2012}}, i.e. when {\it no priors} are applied to the X-ray observables -- mass relation parameters. `Photo-z' and `spectro-z' redshift accuracies are implemented here in terms of the width of the redshift slices the data can be binned into, i.e. of width $\Delta z\,(1+z)$ with $\Delta z = 0.05$ and $\Delta z = 0.01$, respectively. This distinction is to be intended only for the counts experiment, as for the clustering the slicing in redshift-space is limited by shot-noise to a handful of bins in redshift space. The pessimistic and optimistic scenarios which we refer to throughout the paper also differ according to the strength of the priors which can be placed on the observable-mass relation parameters and according to the minimum cluster mass which can be reliably modeled and therefore included in the analysis.}
\caption{\label{TAB:RESULTS_COMPARISON} Comparison among $\eRO$ (optimistic case), {\it Planck} with and without additional data, and the DES results, for cosmological models with constant or evolving DE (see Section \ref{SEC:RESULTS_COMPARISON} for details). The results for the 1st-year constraints from the DES are from \textcolor{blue}{Troxel et al. 2017}. As a future, specific example of StageIV experiment, we give the predictions for the {\it Euclid} mission according to \textcolor{blue}{Giannantonio et al. 2012} (weak-lensing + 2D spectroscopic clustering of galaxies) and \textcolor{blue}{Sartoris et al. 2016} (cluster counts and 3D power spectrum but assuming {\it perfect knowledge} of the cluster mass-observable scaling relations). Notes: (1) {\it Planck} = base\_plikHM\_TTTEEE\_lowTEB;{\it Planck} + BAO + JLA = base-w-wa\_plikHM\_TTTEEE\_lowTEB\_BAO\_H070p6\_JLA. (2) Here, perfect knowledge of the cluster scaling relations is assumed which leads to very optimistic constraints.}
\caption[Results of patient-to-patient transfer and patient-specific fine-tuning.]{Results of patient-to-patient transfer and patient-specific fine-tuning. The $x$-axis indicates number of images used from patient 2. The class IoU is denoted by symbols for \emph{vocal folds} (\tikz\draw[red, thick] (0,0) circle (.5ex);), other tissue (\textcolor{myblue}{$ \mathbf{\times} $}), \emph{glottal space} (\textcolor{darkgreen}{$\mathbf{+}$}), \emph{surgical tool} (\tikz\draw[orange,fill=orange] (0,0) circle (.35ex);), \emph{intubation} (\textcolor{darkyellow}{$ \mathbf{\diamond} $}) and the mean IoU is denoted by black bar (\textbf{--}).}
\caption{Success rates of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a cart pole}
\caption{Comparison of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a cartpole. \textit{Top left: cart position; Top right: pole angle; Botton left: Gaussian noise; Botton right: Poisson noise}.}
\caption{Success rates of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a quadrotor}
\caption{Comparison of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a quadrotor with 3000 trajectories in sampling. \textit{Top left: x position; Top right: y position; Botton left: z position; Botton right: roll angle}.}
\caption{Comparison of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a quadrotor with 3000 trajectories in sampling. \textit{Top left: x velocity; Top right: y velocity; Botton left: z velocity; Botton right: roll rate}.}
\caption{Comparison of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a quadrotor with 6000 trajectories in sampling. \textit{Top left: x position; Top right: y position; Botton left: z position; Botton right: roll angle}.}
\caption{Comparison of \textcolor{blue}{old} and \textcolor{green}{new} algorithm on a quadrotor with 3000 trajectories in sampling. \textit{Top left: x velocity; Top right: y velocity; Botton left: z velocity; Botton right: roll rate}.}
\caption{Comparison of total variance of trajectories resulted from \textcolor{blue}{old} and \textcolor{green}{new} algorithm under high ($\bm{\Sigma}_J$=20) and low ($\bm{\Sigma}_J$=5) levels of jump noise amplitude with 3000 and 6000 trajectories in sampling ($\nu$=0.2).}
\caption{Architectures of the stack of convolutional layers $g_{m}^{c} \left( \; . \; ; \boldsymbol{\phi}_{m}^{c, 0} \right)$ that is applied to the context portion $\mathbf{X}_{0}$ located on the left side of the image block to be predicted ((a) $m = 4$, (c) $m = 16$, and (e) $m = 64$) and the stack of transposed convolutional layers $g_{m}^{t} \left(\; .\; ;\boldsymbol{\phi}_{m}^{t} \right)$ ((b) $m = 4$, (d) $m = 16$, and (f) $m = 64$). \enquote{conv} means convolution and \enquote{tconv} means transposed convolution.}
\caption{Example power labels from Separated-CNN on the Grouped formulation. For \textit{correct} labels, text segments that may signal the power relation are \setulcolor{Green}\ul{underlined in green}; for \textit{incorrect} labels, potentially confusing power signals are \setulcolor{Red}\ul{underlined in red}. (text segments chosen based on our qualitative judgment).}
\caption{Actuator and sensor configuration of the considered experimental setup. The locations of actuators used for control and identification are marked with red crosses (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[tuewarmred,very thick][-] (0,-2.75) -- (5.5,2.75);\protect\draw[tuewarmred,very thick][-] (0,2.75) -- (5.5,-2.75);}) actuators used only for identification as blue crosses (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[tueblue,very thick][-] (0,-2.75) -- (5.5,2.75);\protect\draw[tueblue,very thick][-] (0,2.75) -- (5.5,-2.75);}), sensors used for identification and control as red circles (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[tuewarmred,very thick] (7,0) circle (2.75);}), and sensors used only for identification as green circles (\protect\tikz[baseline=-0.6ex,x=1pt,y=1pt]{\protect\draw[tuelightgreen,very thick] (7,0) circle (2.75);}).}
\caption{{\it Left} panel: The blue dots are the best-fit values for $\sigma_z$ obtained when fitting the curves $\lambda(z)$ for each cluster in the \redmapper\cluster catalogue (see text). The{\it red squares} represent the $5$ percentile of the $\sigma_z$ distribution estimated in redshift bins of width $\Delta z=0.01$. The {\it solid orange} line shows the model for $\sigma_z(z)$ adopted for the analysis. {\it Right} panels: The {\it blue solid} lines are the measured $\lambda(z)$ for the three clusters labeled with triangles in the {\it left} panel; the {\it red solid} lines represent the best-fit model for $w(\Delta z,z_{\rm cl})$; for comparison, the {\it black dashed} lines show the redshift kernel expected for {\it clear} l.o.s. clusters, i.e. assuming the calibrated $\sigma_{z}(z)$ in the computation of the kernel $w(\Delta z,z_{\rm cl})$.}
\caption[]{V5668 Sgr: UV emission line blue-to-red peak ratios \textcolor{blue}{}}
\caption{Comparison of the [Si VI] 1.96$\mu$ profile from day 350 (dash) with He II 1640\AA\from the STIS spectrum (solid). This highlights the difference in the extinction between the infrared and UV. See text for details.}
\caption[]{V5668 Sgr: HST/STIS log of observations. \textcolor{blue}{}}
\caption{\color{red} $\Box$ #1}
\caption{H.E.S.S. data and analysis results. The first column represents the data set. The second and third columns show the number of signal + background events around the source position, and background events from the off-source region, respectively. The fourth column shows the excess in \grays{}. The background normalisation ($\alpha$) is $\approx 0.022$.}
\caption{SED of Cen A \gray{} core. \textit{Fermi}-LAT and H.E.S.S. data points along with a high-energy power-law extrapolation of the $\gamma$-ray spectrum measured below the break energy. Eight years of \textit{Fermi}-LAT data and 213 hours of H.E.S.S. data were used. Statistical error bars are shown.}
\caption{A schematic `map' of all dimension 6 operators allowed by Lorentz symmetry, cut in half about its axis of symmetry $\sum h =0$ (reflected in this line are the hermitian conjugates of the operators shown). Arrows show the movement induced by equation of motion relations and commutation of derivative relations in the space of dim 6 operators, colour-coded to show \textcolor{blue}{$\phi$ EOMs (blue)}, \textcolor{green}{$\psi$ EOMs (green, short dashed)}, \textcolor{red}{$F$ EOMs (red, long dashed)}, and \textcolor{gray}{replacing derivatives with field strengths (grey, dash-dotted)}. \label{fig:moves}}
\caption{Photon mass $m$ measured by the flux correlation $C(\tau)=\langle \theta(\tau_0) \theta(\tau_0+\tau) \rangle\sim \exp(-m\tau)$ with $\theta(\tau) = \sum_{\square} \sin \left( \text{curl} \phi(\tau) \right)$. Zero photon mass is observed in the U1D phase, finite photon mass is observed in the confined phase. Here we plot $N_f=2$ case.} \label{fig:n2fluxcorr} \end{figure} \begin{figure}[htp] \includegraphics[width=\columnwidth]{n2decay.pdf} \caption{The log-log plot of real space decay of (a) spin correlation functions and (b) dimer correlation functions for $N_f=2$ in the U1D phase (at $J=1.25<J_c$). The slope gives a good estimation of the scaling dimension of spin and dimer. Note that to avoid even-odd oscillation in the finite size data, here only the distance $r=$ odd points are plotted in the U1D phase. For other $N_f$ cases in the following, we adopt to the same strategy. } \label{fig:n2decay} \end{figure} A simplest way to detect the deconfinement-confinement transition may be the Wilson loops, but it is known that in the presence of matter field, it cannot be used to detect the topological order of the deconfined phase (the U1D phase), such effects have been discussed in the literature~\cite{hermele2004a,gregor2011}. One suitable way here to demonstrate the deconfinement-confinement transition is to show how the photon mass changes over $J$. As soon as the matter fields bind to form the particle-hole condensate, we expect monopoles to proliferate and to generate a photon mass. The photon mass can be measured by the correlation of flux quantity $\theta(\tau)$~\cite{burkitt1988glueballs}, which is defined as $\theta(\tau) = \sum_{\square} \sin \left(\text{curl} \phi(\tau) \right)$. The photon mass $m$ is related to the correlation of $\theta(\tau)$ by $C(\tau)=\langle \theta(\tau_0) \theta(\tau_0+\tau) \rangle \sim \exp(-m\tau)$. Fig.~\ref{fig:n2fluxcorr} plots the estimated photon mass for different system sizes. we find a signature of absence of photon mass in the U1D phase and a growth of the photon mass in the AFM phase as $J$ increases. However, we want to point out that due to finite size effects, i.e. uncertainties in extracting the exponential decay in $\theta(\tau)$ close to the transition, the estimation of photon mass near phase transition is more qualitative than quantitative. To further understand the properties of the deconfined phase, we measure the real-space correlation functions in the U1D phase (at $J=1.25<J_c$). As shown in Fig.~\ref{fig:n2decay}(a), the spin-spin correlation shows a power-law with the power $2\Delta_S=3.1(4)$ ($\Delta_S$ is the scaling dimension of spin). Interestingly the dimer-dimer correlation function decays with a similar power-law with the power $2\Delta_D=2.9(4)$ ($\Delta_D$ is the scaling dimension of dimer) (Fig.~\ref{fig:n2decay}(b)). This result sheds light on the property of the deconfined phase, which is proposed in Refs.~\cite{Rantner2002,Hermele2005} to correspond to the algebraic spin liquid. It has the unique property that as a deconfined state emerging from competing orders, the correlation functions of these competing orders, such as spin-spin, dimer-dimer and bond-bond, have the the same power-law decay. If the data in Fig.~\ref{fig:n2decay} were deep inside the confined phase, the decay of spin-spin and dimer-dimer correlations will be very different. For example, in the N\'eel phase, spin-spin will decay to a constant value and dimer-dimer correlation will decay exponentially. Therefore, our data in Fig.~\ref{fig:n2decay} provide supporting evidence of the algebraic spin liquid behavior of the U1D in Fig.~\ref{fig:phasediagram} at $N_f=2$. \subsubsection{$N_f=4$} Next we turn to the $N_f=4$ case where we also observe a U1D phase at small $J$. As shown in Fig.~\ref{fig:n4rvbs}(a), we can follow the crossing points of the correlation ratio of the VBS order parameter for different system sizes, so as to extract (see Fig.~\ref{fig:n4rvbs}(b)) $J_{c1} = 1.2(3)$. The data is consistent with a continuous phase transition from the deconfined phase to the VBS phase. Furthermore, the flux energy per plaquette also supports a continuous transition. \begin{figure}[htp!] \includegraphics[width=\columnwidth]{n4rvbs.pdf} \caption{The VBS correlation ratio through U1D to VBS transition at $N_f=4$. Here $\beta=3L$, $\Delta \tau=0.2$. (b) The $1/L$ extrapolation of the crossings estimates the U1D to VBS transition point $J_{c1}=1.2(3)$ for $N_f=4$.} \label{fig:n4rvbs} \end{figure} \begin{figure}[htp] \includegraphics[width=\columnwidth]{n4decay.pdf} \caption{The log-log plot of real space decay of (a) spin correlation functions and (b) dimer correlation functions for $N_f=4$ in the U1D phase (at $J=1.00<J_c$). The slope gives a good estimation of the scaling dimension of spin and dimer.} \label{fig:n4decay} \end{figure} Fig.~\ref{fig:n4decay} depicts the real space decay of the spin-spin and dimer-dimer correlation functions in the U1D phase for different system sizes. Again, they show similar power-law decay, and the power is estimated to be $2\Delta_S=3.6(3)$ for spin-spin correlation and $2\Delta_D=3.5(3)$ for dimer-dimer correlation. This power law is faster than at $N_f=2$, and is hence consistent with the large-$N_f$ prediction~\cite{Rantner2002,Hermele2005,Hermele2007,Xu2007}. Besides the power-law decay of various competing correlation functions, the situation at $N_f=4$ is even more interesting than that at $N_f=2$. As we further increase $J$, we observe another phase transition from VBS to AFM. This phase transition is discussed in detail in Sec.~\ref{sec:afmvbstransition} of the main text. \begin{figure}[t!] \includegraphics[width=\columnwidth]{eta.pdf} \caption{Dimension of spin (gren circles) and dimer (blue squares) in the U1D phase as a function of $N_f$. The dashed red line corresponds to the $1/N_f$ perturbative calculation, $1+\eta=4-\frac{64}{3\pi^2N_f}$, taken from Ref.~\cite{Rantner2002}. Note here $N_f$ corresponds to the number of four-component Dirac fermions.} \label{fig:eta} \end{figure} \subsubsection{Scaling dimension in U1D phase} The U1D phase found in the small $J$ region is expected to have the same scaling dimension for spin ($\Delta_S$) and dimer ($\Delta_D$)~\cite{Hermele2005}. In fact, according to the large-$N_f$ perturbative renormalization group calculation, these correlation functions decay as \begin{equation} \sim r^{-(4-\frac{64}{3\pi^2 N_f})} \end{equation} with correction at $O(\frac{1}{N^2_f})$~\cite{Rantner2002,Hermele2005,Hermele2007,Xu2007}. Notice that in our case, $N_f$ is the number of fermion flavors on the lattice, while in Refs.~\cite{Rantner2002,Hermele2005,Hermele2007,Xu2007}, $N_f$ is the number of two-component Dirac fermions, which is twice of our $N_f$ due to momentum valley degeneracy. We now compare this theoretical expectation to our numerical simulation results. Fig.~\ref{fig:eta} presents a summary plot of the power law we obtain at $N_f=2$, $4$, $6$, and $8$. The results for $N_f=6$ and $N_f=8$ are detailed in Appendix~\ref{sec:nf6and8}. It is remarkable to see that our data perfectly matches the aforementioned $1/N_f$ perturbative expression. \subsubsection{Theory for confinement transition} We find a U1D to AFM phase transition for the $N_f=2$ case and a U1D to VBS phase transition for the $N_f=4$, $6$ and $8$ cases. These phase transitions should belong to the QED$_3$-Gross-Neveu-O(3) or XY transitions~\cite{Janssen2017}, depending upon the order parameters in the confined phases. For example, at least with large enough (but still finite) $N_f$ when the higher order fermion interactions are clearly irrelevant, the transition between U1D to VBS phase can be described by the following action: \begin{eqnarray} \mathcal{S} = \int d^2x d\tau \sum_{j = 1}^{2N_f} \bar{\psi}_j \gamma \cdot (\partial - i a)\psi_j + u \vec{\phi} \cdot \bar{\psi} \vec{\mu} \psi \nonumber\\ + |\partial \vec{\phi}|^2 + r |\vec{\phi}|^2 + g|\vec{\phi}|^4, \end{eqnarray} where $\psi=(\psi_1, \cdots, \psi_{2N_f})^T$ has $2N_f$ components, $\vec{\phi}$ is a $O(2)$ vector in which the VBS order parameter is embedded. $\vec{\mu} = (\mu^x, \mu^y)$ are two $2N_f \times 2N_f$ matrices in the fermion flavor space. $\bar{\psi}\mu^x\psi$, $\bar{\psi}\mu^y\psi$ are two fermion mass operators that correspond to the VBS in $x$ and $y$ directions respectively. When $r > 0$, $\vec{\phi}$ is gapped out, and the system is in the U1D phase due to the screening of massless fermions to the gauge field. When $r < 0$, $\vec{\phi}$ condenses, the fermions are gapped out, then the compact gauge field is in the confined phase. In the U1D phase, the VBS and antiferromagnetic order parameters should have the same scaling dimension, due to the enlarged SU($2N_f$) symmetry in the low energy field theory, consistent with our data in Fig.~\ref{fig:n8decay} and Figs.~\ref{fig:n2decay},\ref{fig:n4decay} and \ref{fig:n6decay} in Appendix~\ref{sec:phaseboundaries}; while at the critical point $r = 0$, these two order parameters still have power-law correlation functions, but with different scaling dimensions, due to the loss of the SU($2N_f$) symmetry in the infrared. \subsubsection{AFM-VBS transition at $N_f = 4$} \label{sec:afmvbstransition} The situation at $N_f=4$ is even more interesting. As we further increase $J$, we observe another quantum phase transition from VBS to AFM phase. This transition is consistently revealed in three steps in Fig.~\ref{fig:n4rafm}. Fig.~\ref{fig:n4rafm} (a) shows the $r_{\text{AFM}}$ correlation ratio and clearly there is a crossing point signifying the establishment of the AFM long range order. The inset shows the $1/L$ extrapolation of the crossing point and gives rise to $J_{c2}=18(3)$. Fig.~\ref{fig:n4rafm} (b) is the $r_{\text{VBS}}$ correlation ratio and the crossing point in it signifies the vanishing of the VBS order. Inset of Fig.~\ref{fig:n4rafm} (b) gives rise to $J_{c2}=19(5)$, consistent with the onset of the AFM order in Fig.~\ref{fig:n4rafm} (a). \begin{figure}[h!] \includegraphics[width=\columnwidth]{n4largejcombine.pdf} \caption{(a) Antiferromagnetic correlation ratio $r_{\text{AFM}}$ for $N_f=4$. Here $\beta=3L$, $\Delta \tau=0.2$. Insets shows the $1/L$ extrapolation of the crossing point in $r_{\text{AFM}}$ and $J_{c2}=18(3)$. (b) VBS correlation ratio $r_{\text{VBS}}$ for $N_f=4$. The inset is the $1/L$ extrapolation of the crossing point in $r_{\text{VBS}}$ and $J_{c2}=19(5)$. This is consistent with $J_{c2}$ obtained from $r_{\text{AFM}}$ in (a). (c) AFM/VBS correlation ratio $r_{\text{AFM/VBS}}$ for $N_f=4$. The inset is the $1/L$ extrapolation of the crossing point in $r_{\text{AFM/VBS}}$ and $J_{c2}=17(4)$. } \label{fig:n4rafm} \end{figure} The transition from VBS to AFM deserves more attention, apparently the data in Fig.~\ref{fig:n4rafm} suggest a continuous transition, and if it were the case, there is then the possibility that the critical point will acquire larger symmetry group than that in the model in Eq.~\eqref{eq:hamiltonian}. In the case of $Z_2$ gauge field coupled to fermion, as shown in Refs.~\cite{Assaad2016,Gazit2018}, two similar situations with emergent continuous symmetry are also investigated. In the first case, it is at $N_f=3$ that a continuous VBS to AFM phase transition occurs~\cite{Assaad2016}, and in the second case, it is the deconfine-confine phase transition itself at $N_f=2$~\cite{Gazit2018}. Our phase transition at $J_{c2}$ is closer to the former. To further understand the nature of transition from VBS to AFM, we plot the ratio of AFM structure factor and VBS structure factor, \begin{equation} r_{\text{AFM/VBS}}= \frac{\chi_S(\vec{X})}{\chi_D(\vec{M})}. \end{equation} The results are shown in Fig.~\ref{fig:n4rafm} (c), indeed there is a crossing point in the $r_{\text{AFM/VBS}}$, and the $1/L$ extrapolation in the inset of Fig.~\ref{fig:n4rafm} (c) gives rise to $J_{c2}=17(4)$, very consistent with the $J_{c2}$ obtained from the crossings of $r_{\text{AFM}}$ in Fig.~\ref{fig:n4rafm} (a) and $r_{\text{VBS}}$ in Fig.~\ref{fig:n4rafm} (b). This transition is similar to the AFM-to-VBS transition in the $SU(4)$ J-Q model~\cite{Lou2009}, where a continous transition is observed. The transition in that case can be described by a NCCP$^{N-1}$ description with $N=4$~\cite{Kaul2012}, and it is shown that the monopoles are irrelevant at this fix point~\cite{Block2013} and therefore a deconfined quantum critical point~\cite{Senthil2004} is realized. However, in Ref.~\cite{Kaul2012}, on sublattice A of the lattice there is a fundamental representation of SU(4), while on sublattice B there is a anti-fundamental representation, which is different from our case. In our case, with $N_f = 4$ there is effectively a self-conjugate representation of the SU(4) group on every site, thus the field theory for the AFM-to-VBS transition is different from Ref.~\cite{Kaul2012}. According to Ref.~\cite{sachdevread}, the AFM N\'{e}el order in this case has the following grassmanian ground state manifold $\mathcal{M}$: \begin{equation} \mathcal{M} = \frac{U(4)}{U(2)\times U(2)}. \end{equation} To describe this antiferromagnetic state, one can either introduce $N_f = 4$ flavors of fermionic spinons with half-filling, or introduce two color species of bosonic spinons $z_{\alpha,a}$ ($\alpha = 1 ,\cdots 4$, $a = 1,2$), and couple them to a U(2) gauge field (to describe the simplest N\'{e}el order of SU(2) spins, we only need one two component of bosonic spinon coupled with a U(2) gauge field, as Ref.~\cite{Senthil2004}). The U(2) gauge constraint will guarantee that on every site there are fixed number of spinons, and the color space is fully antisymmetric, thus on every site the SU(4) spin is automatically in an antisymmetric self-conjugate representation. Then the field theory for the N\'{e}el-VBS transition is \begin{equation} \mathcal{S} = \int d^2x d\tau \|(\partial - i a - i \sum_{l = 1\cdots 3} a^l \tau^l)z|^2 + r |z|^2 + \cdots \end{equation} where $a_\mu$ and $a^l_{\mu}$ are gauge fields corresponding to the U(1) and SU(2) subgroups of U(2). Note that these gauge fields are ``emergent" gauge fields, which are different from the explicit gauge field in our original simulation. When $r < 0$, $z_{\alpha,a}$ condenses, and leads to the antiferromagnetic state with ground state manifold $\frac{U(4)}{U(2)\times U(2)}$. When $r > 0$, $z_{\alpha,a}$ is gapped out, and the gauge fields will be confined. Here we assume that the U(1) compact gauge field $a_\mu$ still has the quadru-monopole proliferation, which leads to the VBS phase like the original deconfined QCP theory for the SU(2) spins~\cite{Senthil2004}. One of the crucial properties of the deconfined QCP is the ``intertwinement" between order parameters on two sides of the transition, which can be captured by a topological term which treats the N\'{e}el and VBS order parameters on an equal footing~\cite{senthilfisher}. In the current case with SU(4) spin symmetry, one can also introduce a topological term that captures the ``intertwinement" between the the SU(4) N\'{e}el and VBS orders with a topological term. To do this, we need to embed both the N\'{e}el and VBS order parameters into a larger manifold. One way to parameterize the N\'{e}el order manifold is $\mathcal{N} = U^\dagger \Omega U$, where $\Omega$ is a $4\times 4$ diagonal matrix $\Omega = \mathrm{diag}(\mathbf{1}_{2\times 2}, - \mathbf{1}_{2\times 2})$, and $U$ is a SU(4) matrix. $\mathcal{N}$ is a $4\times 4$ Hermitian matrix with constraint $ \mathcal{N}^2 = 1$. Now we introduce a larger $8\times 8$ matrix $\mathcal{P}$ which includes both the N\'{e}el and VBS order parameters: \begin{equation} \mathcal{P} = \cos(\theta) \mathcal{N} \otimes \tau^z + \sin(\theta) \mathbf{1}_{4\times 4} \otimes \left(V_x \tau^x + V_y \tau^y \right), \end{equation} where $(V_x, V_y)$ is a two component order parameter for the VBS phase, and $V_x^2 + V_y^2 = 1$. The order parameter $\mathcal{P}$ unifies the SU(4) N\'{e}el and VBS order parameters, just like the O(5) vector order parameter introduced in Ref.~\cite{senthilfisher}. The topological term that captures the ``intertwinement" between N\'{e}el and VBS order parameters is a Wess-Zumino-Witten term: \begin{equation} \mathcal{S}_{wzw} \sim \int d^2x d\tau \int_0^1 du \\epsilon_{\mu\nu\rho\sigma}\mathrm{tr}[\mathcal{P} \partial_\mu \mathcal{P} \partial_\nu \mathcal{P} \partial_\rho \mathcal{P} \partial_\sigma \mathcal{P}]. \label{wzw} \end{equation} Using the same technique in Ref.~\cite{groversenthil}, one can show that at the vortex core of the VBS order parameter, there is a spinon with self-conjugate representation, which is consistent with intuition. In fact, the O(5) WZW term introduce in Ref.~\cite{senthilfisher} can be written in the same form as Eq.~\eqref{wzw}, as long as we replace $\mathcal{P}$ in Eq.~\eqref{wzw} by a $4\times 4$ Hermitian matrix order parameter $\mathcal{P} = \vec{n} \cdot \vec{\Gamma}$, where $\vec{\Gamma}$ are five Gamma matrices, and $\vec{n}$ is the O(5) vector introduced in Ref.~\cite{senthilfisher}. This topological term can be viewed as the low-energy effective field theory of the $\pi-$flux state of the SU(4) antiferromagnet, which again is described by a QED$_3$ with eight flavors of Dirac fermions~\cite{Hermele2005}, but again the gauge field of this QED$_3$ is an emergent gauge field which is different from the gauge field introduced in the original model that we simulate. The WZW term Eq.~\eqref{wzw} can be derived by coupling the $8\times 8$ matrix order parameter $\mathcal{P}$ to the eight flavors of Dirac fermions of the $\pi-$flux state, and integrate out the fermions following the standard procedure of Ref.~\cite{abanov}. Our data, the crossing of $r_{\text{AFM}/\text{VBS}}$ in Fig.~\ref{fig:n4rafm} (c), suggest that the AFM and VBS order parameters have the same scaling dimension at this transition, which is consistent with the emergent SU(8) symmetry of the $\pi-$flux state of the SU(4) antiferromagnet. The large SU(8) symmetry, if indeed exists at the AFM-VBS transition, will ensure that many other order parameters have the same scaling dimension as AFM and VBS order parameters~\cite{Hermele2005}. These order parameters will also have similar fractionalization dynamical signatures in their spectral functions as AFM and VBS order parameters. \section{Conclusions} \label{sec:discuss} Using large scale DQMC, we investigated the compact $U(1)$ gauge field theory coupled to Dirac fermion matter fields in $(2+1)$D and variable flavor number $N_f$: i.e. cQED$_3$. With our simulations we mapped out the entire ground state phase diagram in the flavor $N_f$ and gauge field fluctuation $J$ strength plane. Our results are summarized in Fig.~\ref{fig:phasediagram}(a). Most importantly, signatures supporting stable U(1) deconfined (U1D) phases were discovered at $N_f=8$ and $6$, and evidence of the U1D phase at $N_f=4$ and $2$ were also found. The properties of the deconfined phase are consistent with the proposal of algebraic spin liquid, in which, various competing orders (AFM order and VBS order, for example) all have algebraic correlation with identical power-laws in real-space. The decay power is found to quantitatively converge to the large-$N_f$ predictions~\cite{Rantner2002,Hermele2005,Hermele2007,Xu2007}. The transition between the deconfined and confined phases at various $N_f$ were determined using the RG invariant correlation ratios. At $N_f=2$ the transition occurs between the U1D and AFM phases. Since the AFM corresponds to O(3) symmetry breaking, the critical theory should be described by the QED$_3$-Gross-Neveu O(3) universality class. In contrast, at larger values of $N_f$ the ordered phases corresponds to VBS. The dynamical generation of the two VBS mass terms is described by the QED$_3$-Gross-Neveu O(2) universality class. As far as we know, the QED$_3$-Gross-Neveu O(2) or O(3) transition have not been investigated numerically before in an unbiased simulation. It is certainly worthwhile to further carefully study the critical properties of these transitions via QMC simulations and compare with future analytical calculations. In particular, the QED$_3$-Gross-Neveu Ising transition has been investigated recently with perturbative RG calculations~\cite{Janssen2017,Ihrig2018}. Aside from the QED$_3$-Gross-Neveu transitions, we have found evidence for a direct and continuous transition between the AFM and VBS states in the confined region of the phase diagram at $N_f = 4$. Since we have on average two fermions per site, we should consider the antisymmetric self-conjugate representation of the SU(4) group. We have presented various theoretical descriptions of this putative deconfined quantum phase transition in terms of multi-flavored spinons coupled to emergent U(1) and SU(2) gauge fields. We also discussed the effective low energy field theory with a topological term that captures the intertwinement between the magnetic and VBS orders, and its connection to the $\pi-$flux state of the SU(4) spin system discussed in Ref.~\cite{Hermele2005}. Numerical support for emergent symmetries was provided. In the future, measurements of the conserved current operators related with such emergent continuous symmetries~\cite{Ma2018Current} can be performed. Finally, the confining phase transitions in our model, as well as the possible deconfined quantum critical point at $N_f=4$, will have distinct and very interesting dynamical properties in their spectral functions, that can be further explored in QMC simulations. Such calculations provide experimentally accessible signatures of exotic states of matter where emergent gauge fields, fractionalized excitations, can be traced. Similar attempts have recently been applied to the deconfined quantum critical point in pure spin model~\cite{Ma2018}, emergent $Z_2$ spin liquid at (2+1)D~\cite{GYSun2018} and $U(1)$ spin liquid at (3+1)D~\cite{Huang2018} and the $Z_2$ counterpart of our model~\cite{Assaad2016}. In the present cQED$_3$ model, dynamical measurements in the QMC simulation plus state-of-art analytical continuation~\cite{Sandvik2015,Shao2017b,GYSun2018} can help to reveal more fundamental physical understanding of these exotic quantum phase transitions. \begin{acknowledgements} The authors thank Lukas Janssen, Ribhu Kaul, Steven Kivelson, Thomas Lang, Srinivas Raghu, Subir Sachdev, Michael Scherer, Yi-Zhuang You, Ashvin Vishwanath, Chong Wang for helpful discussions. FFA would like to thank T. Grover and Zhenjiu Wang for discussions on related projects. XYX, YQ, LZ, CX acknowledge the hospitality of the International Collaboration Center at Institute of Physics, Chinese Academy of Sciences, this work is finalized during the program "International Workshop on New Paradigms in Quantum Matter" at the Center. XYX is thankful for the support of HKRGC through through C6026-16W, 16324216 and 16307117. YQ acknowledges support from Minstry of Science and Technology of China under grant numbers 2015CB921700, and from National Science Foundation of China under grant number 11874115. ZYM acknowledges supports from the National Key R\&D Program (2016YFA0300502), the Strategic Priority Research Program of CAS (XDB28000000), and the National Science Foundation of China (11574359, 11674370). FFA thanks the DFG research unit FOR1807 for financial support. LZ is supported by the National Key R\&D Program of China (Grant No. 2018YFA0305802) and the Key Research Program of Chinese Academy of Sciences (Nos. XDPB08-4 and XDB28040200). We thank the Center for Quantum Simulation Sciences in the Institute of Physics, Chinese Academy of Sciences, the Tianhe-1A platform at the National Supercomputer Center in Tianjin and Tianhe-2 platform at the National Supercomputer Center in Guangzhou for their technical support and generous allocation of CPU time.\end{acknowledgements} \appendix \section{Connection to high energy lattice cQED$_3$ action} As we discussed in the Sec.~\ref{sec:model}, after the path integral of the rotor degrees of freedom in a rotor model with fermion in Eq.~\eqref{eq:hamiltonian}, the action of cQED$_3$ coupled to fermionic matter is obtained explicitly. In the high energy lattice cQED$_3$ action, the Lagrangian for pure gauge field part takes the form \begin{equation} L_\phi = K_{\tau} \sum_{\langle ij \rangle} \left(\cos(\phi_{ij}(\tau+1)-\phi_{ij}(\tau)) \right)+ K_{r} \sum_{\square}\cos \left(\text{curl} \phi \right)\label{eq:hep-lagrangian} \end{equation} where $K_{\tau}<0$ and $K_r<0$ with $|K_{\tau}|=|K_r|$. Compared with the Lagrangian defined in Eq.~\eqref{eq:lagrangian}, we study the case $K_{\tau}<0$ and $K_r>0$ with $|K_{\tau}|\ne |K_r|$. As we can always rescale space and time to restore the Lorentz symmetry, the difference between $|K_{\tau}| = |K_r|$ and $|K_{\tau} \ne |K_r|$ is trivial. Actually, our model can be exactly mapped to the case of $K_r < 0$ and the fermion hopping with a staggered phase factor as follows: \[\phi_{i,i+\hat x}\rightarrow\phi_{i,i+\hat x}+m_y(i)\pi,\] where $m_y(i)$ is 1 (0) if the $y$-coordinate of $i$ is odd (even), respectively. Therefore, our convention is equivalent to the high energy lattice cQED$_3$ action. As we mentioned in the main text, the Dirac fermion in our model is realized, because the $K_r$ term prefers $\pi$-flux through each plaquette, and the $\pi$-flux doubles the unit cell. Following the standard literature such as Ref.~\cite{Hermele2005}, if we start with $N_f$ flavors of one-component fermions on the lattice (like Eq.~\eqref{eq:hamiltonian}), at low energy there will be $2N_f$ flavors of 2-component Dirac fermions. \section{Pseudo-unitary group $SU(n,m)$ and the absence of sign problem} \label{sec:psu} As we discussed in the main text, the fermion determinant for one flavor is $\det \big( I_{n+m}+\prod_{z=1}^{L_\tau}B_z \big)$, where $n$ and $m$ are the numbers of sites in the two sublattices, and $I_{n+m}$ denotes the $(n+m)$-dimensional identity matrix. $B_{z}=e^{h_{z}}$, where $h_{z}$ has the following structure, \begin{equation} h_{z}= \begin{pmatrix} 0_{n} & T_{z} \\ T_{z}^{\dag} & 0_{m} \end{pmatrix}, \end{equation} and $T_{z}$ is the hopping matrix between different sublattices. $B_{z}$ matrices satisfy (1) $B_{z}^{\dag}\eta B_{z}=\eta$, where $\eta=\mathrm{diag}(I_{n},-I_{m})$, and (2) $\mathrm{det}B_{z}=1$, thus their products generate the pseudo-unitary group $\mathrm{SU}(n,m)$. \emph{Theorem.} For any $D\in \mathrm{SU}(n,m)$, $\det (I_{n+m}+D)\in \mathbb{R}$. \emph{Proof.} First, suppose that $\lambda$ is an eigenvalue of $D$, $Dv=\lambda v$, then $D^{\dag}\eta v=\lambda^{-1}\eta v$, and $D^{T}\eta v^{*}=(\lambda^{*})^{-1}\eta v^{*}$, hence $(\lambda^{*})^{-1}$ is an eigenvalue of $D^{T}$, and is thus also an eigenvalue of $D$. Denote the eigenvalues of $D$ by $\lambda_{i}$, $1\leq i\leq n+m$, then $\det (I_{n+m}+D)=\prod_{i}(1+\lambda_{i})$. We then treat the eigenvalues on the unit circle and those not on the unit circle separately. For those not on the unit circle, \begin{equation} \prod_{i,|\lambda_{i}|\ne1}(1+\lambda_{i})=\prod_{i,|\lambda_{i}|<1}(1+\lambda_{i})(1+(\lambda_{i}^{*})^{-1})=\prod_{i,|\lambda_{i}|<1}\frac{|1+\lambda_{i}|^{2}}{\lambda_{i}^{*}}. \end{equation} For those on the unit circle, denoting $\lambda_i = e^{i\theta_i}$, $-\pi<\theta_i\leq \pi$, we have \begin{equation} \prod_{i,|\lambda_{i}|=1}(1+\lambda_{i})=\prod_{i,|\lambda_{i}|=1}2\cos(\theta_{i}/2)e^{i\theta_{i}/2}. \end{equation} Therefore, we find \begin{equation} \begin{split} (\det(I_{n+m}+D))^{2}&=|\det(I_{n+m}+D)|^{2}\prod_{i}\frac{\lambda_{i}}{|\lambda_{i}|}\\ &=|\det(I_{n+m}+D)|^{2}> 0, \end{split} \end{equation} hence $\det(I_{n+m}+D)\in \mathbb{R}$. This theorem implies that for an even number of flavors fermions, the model is free of sign problem for any hopping matrices $T_{z}$. For instance, this is true for both abelian and nonabelian gauge fields. It is easy to find an example \begin{equation} D= \begin{pmatrix} -\sqrt{2}& 1\\ 1& -\sqrt{2} \end{pmatrix} \in\mathrm{SU}(n,m) \end{equation} such that $\det (I+D)<0$. Therefore, the absence of sign problem does not hold for an odd number of fermions in general. However, it does hold for models with fermions coupled to $\mathbb{Z}_{2}$ gauge fields \cite{Assaad2016, Gazit2017, Gazit2018}. \section{Confinement transition for $N_f=6$ and $N_f=8$} \label{sec:nf6and8} In this section we discuss the results for $N_f=6$ and $N_f=8$. \begin{figure}[htp] \includegraphics[width=\columnwidth]{n6rvbs.pdf} \caption{The VBS correlation ratio through U1D to VBS transition at $N_f=6$. Here $\beta=2L$, $\Delta \tau=0.1$. (b) The $1/L$ extrapolation of the crossings estimates the U1D to VBS transition point at $J_c=1.9(3)$ for $N_f=6$.} \label{fig:n6rvbs} \end{figure} \begin{figure}[htp] \includegraphics[width=\columnwidth]{n6decay.pdf} \caption{The log-log plot of real space decay of (a) spin correlation functions and (b) dimer correlation functions for $N_f=6$ in the U1D phase (at $J=1.40<J_c$). The slope gives a good estimation of the scaling dimension of spin and dimer.} \label{fig:n6decay} \end{figure} As shown in Fig.~\ref{fig:n6rvbs} and Fig.~\ref{fig:n8rvbs}, and the corresponding insets, we estimate $J_c=1.9(3)$ for $N_f=6$ and $J_c=2.5(1)$ for $N_f=8$. Again, the data are consistent with a continuous transition between the deconfined UID and confined VBS phases. Correspondingly, the flux energy per plaquette behaves as smooth function across the critical point. \begin{figure}[h!] \includegraphics[width=\columnwidth]{n8decay.pdf} \caption{The log-log plot of real space decay of (a) spin correlation functions and (b) dimer correlation functions for $N_f=8$ in the U1D phase (at $J=2.00<J_c$). The slope gives a good estimation of the scaling dimension of spin and dimer.} \label{fig:n8decay} \end{figure} \begin{figure}[h!] \includegraphics[width=\columnwidth]{n8rvbs.pdf} \caption{The VBS correlation ratio through U1D to VBS transition at $N_f=8$. Here $\beta=2L$, $\Delta \tau=0.1$. (b) The $1/L$ extrapolation of the crossings estimates U1D to VBS transition point at $J_c=2.5(1)$ for $N_f=8$.} \label{fig:n8rvbs} \end{figure} More interestingly, we plot the spin-spin and dimer-dimer correlation function in real-space for $N_f=6$ at $J=1.4$ in Fig.~\ref{fig:n6decay} and for $N_f=8$ at $J=2.0$ in Fig.~\ref{fig:n8decay}, respectively. In Fig.~\ref{fig:n6decay}, the spin-spin and dimer-dimer correlation functions show the similar power-law decay with $2\Delta_S=3.8(2)$ and $2\Delta_D=3.6(3)$. In Fig.~\ref{fig:n8decay}, both correlation functions decay with similar power, with $2\Delta_S=3.8(2)$ and $2\Delta_D=3.4(5)$. On the whole our data provides concrete evidence that the deconfined phase in our model at various values of $N_f$ belong to the algebraic spin liquid~\cite{Rantner2002,Hermele2005,Hermele2007,Xu2007}. One can foresee that with a further increase of $N_f$, we will reach the expected power-law behavior $\sim r^{-4}$. \section{Flux energy per plaquette} \label{sec:fluxenergy} \begin{figure}[thp!] \includegraphics[width=\columnwidth]{n2fluxenergy.pdf} \caption{Flux energy per plaquette along the same $J$ path. There is no sigularity around $J_c$, suggesting it is a continuous phase transition.} \label{fig:n2fluxenergy} \end{figure} To characterize the continuous nature of confined and deconfined phase transition, we also measured the flux energy per plaquette $\left\langle\frac{1}{L^2} \sum_{\square}\cos \left(\text{curl} \hat{\theta} \right)\right\rangle$. Fig.~\ref{fig:n2fluxenergy} depicts our result at $N_f=2$. For other $N_f$'s, the flux energy per plaquette has a similar continuous behavior. \begin{figure}[htp] \includegraphics[width=\columnwidth]{n2QQ00.pdf} \caption{Uniform structure factor of $\hat{Q}_i$ (defined in Eq.~\ref{eq:cq}) for $N_f=2$. Note that $\beta=4L$ here. For all $J$s, this uniform structure factor extrapolated to zero in thermodynamic limit. Thus, the constrain of $\hat{Q}_i=0$ is dynamically enforced. } \label{fig:n2QQ00} \end{figure} \section{Dynamically generated constraint} As mentioned in the main text, our model corresponds to an unconstrained gauge theory. As such, the Gauss law will be dynamically imposed and $\hat{Q}_i$, defined in Eq.~\ref{eq:Qi}, should converge to constant value in the zero temperature limit. We studied the uniform structure factor of $\hat{Q}_i$ by calculating \begin{equation} C_{Q} = \frac{1}{L^2} \sum_{ij} \langle \hat{Q}_i \hat{Q}_j\rangle. \label{eq:cq} \end{equation} We find that the uniform structure factor of $\hat{Q}_i$ defined above extrapolates to zero in thermodynamic limit as showed in Fig.~\ref{fig:n2QQ00} for $N_f=2$. Other $N_f$ cases show similar behavior. \begin{figure}[htp] \includegraphics[width=\columnwidth]{accep.pdf} \caption{Local update acceptance ratio at $N_f=2$ with $L=6$. Note the acceptance ratio for other sizes is almost same, not shown here.} \label{fig:accep} \end{figure} \section{Performance of DQMC on cQED$_3$ coupled to fermionic matter} As we discussed in the main text, in the DQMC simulation, we use local updates, which flip the gauge variables $\phi_b(\tau)$ ($\in[0,2\pi)$) on space-time lattice one by one and we call one scan of the whole space-time lattice as one sweep, which is usually called one Monte Carlo step in DQMC. For cQED$_3$ problem, we designed a specific fast update method, which greatly improves the computation efficiency, more accurately, by making fast update still work here thus reducing huge time cost for each sweep. As a first attempt to study this challenging problem of cQED$_3$ coupled to fermionic matter in condensed matter field by DQMC method with local update strategy, how well does it work here needs a demonstration. Following is a detailed discussion of the performance of the method. First important quantity associated with the efficiency of the method is the acceptance ratio. Fig.~\ref{fig:accep} illustrated the acceptance ratio for different $J$ at $N_f=2$. The acceptance ratio reduces as $J$ becomes smaller. Fortunately, the acceptance ratio deep in the U1D phase is still quite large, for example at $J=0.75 < Jc=1.6(2)$, the acceptance ratio is $\sim 10\%$. \begin{figure}[htb] \includegraphics[width=\columnwidth]{flux_sweep.pdf} \caption{Net flux sweep serials in time slice plane $\tau$ and $\tau'$ at $N_f=2$ with $L=12$ (a) inside U1D phase and (b) inside AFM phase. Here $\tau=\Delta\tau$ and $\tau'=8\Delta\tau$. The flux sweep serials are plotted in the interval of 20 sweeps. And a $L^2/2$ is added to shift it to be centred around zero. } \label{fig:flux_sweep} \end{figure} Second important quantity which reflects the efficiency of our method to the specific problem we studied is how quickly does the net flux change in each time plane with Monte Carlo steps. Flux in each plaquette can be written as $\sum_{b\in \square} \phi_b = \Phi_{\square} + 2\pi m_{\square}$ with $\Phi_{\square} \in [0,2\pi)$ and $m_{\square}$ an integer. The net flux in one time slice plane $M(\tau)$ is defined as a sum of $m_{\square}$ of each plaquette in the time slice plane $\tau$, $M(\tau)=\sum_{\square} m_{\square}(\tau)$. Fig.~\ref{fig:flux_sweep} showed such net flux sweep series both inside U1D phase (Fig.~\ref{fig:flux_sweep}(a)) and inside AFM phase (Fig.~\ref{fig:flux_sweep}(b)) at $N_f=2$ with $L=12$ at different time slices $\tau$ and $\tau'$. In the U1D phase, it favors $\pi$($-\pi$)-flux in each plaquette, and the net flux in each time slice plane seldom changes and is weakly fluctuating between different time slices, while in the AFM phase, the net flux changes almost randomly with more extended values and large fluctuations between different time slices plane, as a consequence of proliferate of monopoles. \bibliography{main_full} \end{document} }
\caption{The height between the bottom of the droplet and the surface, $\Delta z$, as function of time, $t$, for silicone oil at reduced ambient pressure (\protect\includegraphics{sil001}), silicone oil at normal ambient pressure (\protect\includegraphics{sil100}), ethanol at reduced ambient pressure (\protect\includegraphics{eth001}), and ethanol at normal ambient pressure (\protect\includegraphics{eth100}). Impact time, $t = 0$, is defined as the moment the droplet would have hit the surface if no gas film formed under the droplet. This trajectory is described by the line $\Delta z/D = - t/\tau$. The inset shows the deviation from the impact trajectory due to gas film formation.}
\caption{% The insert in frame (a) shows a schematic drawing of the parameters referred to in the main text. $r_{\txt{c}}$, and $z_{\txt{c}}$ are the radial position and height, respectively, of the cusp which forms as the droplet spreads over the surface. $r_{\txt{cl}}$ is the radial position of the contact line, and $r_{\txt{l}}$, and $z_{\txt{l}}$ show the radial position and height of the lamella, respectively. (a) The radial position of the contact line, $r_{\txt{cl}}$, as function of time for (\protect\includegraphics{sil001}) silicone oil at reduced ambient pressure, (\protect\includegraphics{sil100}) silicone oil at atmospheric ambient pressure, (\protect\includegraphics{eth001}) ethanol at reduced ambient pressure, and (\protect\includegraphics{eth100}) ethanol at atmospheric ambient pressure. (b) The radial position of the cusp, $r_{\txt{c}}$, as function of time. (c) The height of the lamella/liquid sheet, $z_{\txt{l}}$, as function of time using inertial scaling. (d) The height of the lamella/liquid sheet, $z_{\txt{l}}$, as function of time using viscous scaling. (e) The radial position of the lamella, $r_{\txt{l}}$, as function of time using inertial scaling. (f) The radial position of the lamella, $r_{\txt{l}}$, as function of time using mixed scaling. The theoretical curves in both image (a) and (b) are proposed by \citet{mongruel2009}, and show that both length scales scale with inertia. When using inertial scaling in image (c) no collapse of the data can be observed. However, When viscous scaling is applied in image (d) at early times the lamella thickness for ethanol and silicone oil becomes of the same order. When applying inertial scaling in image (e) a large time delay can be observed between the moment that the lamella can first be detected in the case of the low viscosity ethanol and the high viscosity silicone oil. When applying the scaling suggested by \citet{mongruel2009} in image (f) this time delay is greatly reduced. The error bars in this figure are smaller than the symbols.}
\caption{(a) The volume occupied by the liquid in the liquid sheet for (\protect\includegraphics{sil001}) silicone oil at low ambient pressure, (\protect\includegraphics{sil100}) silicone oil at normal ambient pressure, (\protect\includegraphics{eth001}) ethanol at low ambient pressure, and (\protect\includegraphics{eth100}) ethanol at normal ambient pressure. The influx of liquid for the silicone oil liquid sheet at atmospheric ambient pressure is initially larger than at reduced ambient pressure,. However, later in time a significant reduction can be observed. (b) The surface area covered on the wall by the liquid sheet. Because the contact line moves faster at reduced ambient pressure the surface area covered by the liquid sheet is smaller for both ethanol and silicone oil. (c) The average height of the liquid sheet drops significantly for silicone oil at atmospheric ambient pressure. This causes the liquid sheet to break up. These dynamics are not present for ethanol.}
\caption{The typical dependencies of in-phase (left) and quadrature (right) velocity components measured during up ($\blacksquare$) and down (\textcolor{red}{$\medbullet$}) frequency sweeps in the case of higher voltage excitations, a clear sign of hysteresis can be seen. The solid lines represents the f\mbox{}its based on the model of Duf\mbox{}f\mbox{}ing oscillator (see text for details).}
\caption{Overview of the proposed generative model, encompassing the \textcolor{ctxcol}{\bf context model}, \textcolor{idtcol}{\bf identity model}, \textcolor{faccol}{\bf face model} and \textcolor{labcol}{\bf label model}. Unfilled nodes represent latent variables, shaded nodes are observed, the half-shaded node is observed only for a subset of the indices and uncircled nodes denote fixed hyperparameters. $\bpi_0$ and $(\bpi_c)_{c=1}^C$ are the global and context-wise identity probabilities, $\bomega$ denotes the context probabilities, $(c^*_m)_{m=1}^M$ are the frame-wise context labels, indexed by the frame numbers $(f_n)_{n=1}^N$, $(z_n)_{n=1}^N$ are the latent identity indicators, $(\*x_n)_{n=1}^N$ are the face observations and $(y_n)_{n=1}^N$ are the respective name annotations, $(\theta^*_i)_{i=1}^\infty$ are the parameters of the face model and $(y^*_i)_{i=1}^\infty$ are the identities' name labels. See text for descriptions of the remaining symbols.}
\caption{Identity clustering consistency. Markers on the horizontal axis (\protect\tikz{\protect\node[star,fill=red!80!black,star point ratio=2.25,inner sep=1pt]{};}) indicate when new people are met for the first time.}
\caption{A series of analytical ISFs at $\Delta K=1.0\,${\AA}$^{-1}$ given by (\ref{eqn:ISFProductForm}), \textcolor{blue}{associated with the GLE with the friction kernel \ref{eqn:expKernel}, where the overall strength of friction is $\gamma=2.0\,$ps$^{-1}$ but different values of the cutoff frequency $\omega_{c}$ are explored.} The ISFs have been normalised to each other at long times, where they all coincide as shown in the inset. At increasingly small values of $\omega_{c}$, the relative amplitude of the fast decay at shorter times increases in accordance with the factor in (\ref{eqn:initialDropSize}). \textcolor{red}{The curve labelled $\omega_{c}=\infty$ is the Langevin result $I_{L}(\Delta K,t)$, given by (\ref{eqn:ISFLangevinResult}).}}
\caption{Simulated ISF (red points) on the flat surface, at $\Delta K=1.0\,${\AA}$^{-1}$. The key simulation parameters are $\gamma=2.0\,$ps$^{-1}$, $\omega_{c}=1.0\,$ps$^{-1}$, such that $\omega_{c}<\gamma$, namely we are well into the regime where memory is important. The analytical result is overlaid (solid red curves) and shows good agreement with the simulation results. The inset demonstrates the agreement of the simulated and analytical results in the regime where the ISF decays exponentially to zero. Additional simulation parameters are specified in the main text. \textcolor{red}{The analogous results for a Langevin simulation at the same value of $\gamma$ are shown in blue. The emergence of a large ratio between the results of the two simulations can be seen clearly in the inset.}}
\caption{Hand and object detection on GTEA Dataset. Hand probability map (\textcolor{red}{red}) is overlapping with the raw image. Hand bounding box (\textcolor{red}{red}) and object bounding box (\textcolor{green}{green}) are draw on top of the image. Examples are selected from 10 different actions (Top 5: close, fold, open, pour, put; Bottom 5: scoop, shape, spread, stir, take).}
\caption{\textbf{Top}: A brief pipeline of MMAN. Two discriminators are attached to a CNN-based generator ($G$). The Macro $D$ works on the low-resolution label map and has a global receptive field, focusing on semantic consistency. Micro $D$ focuses on multiple patches and has small receptive fields on high-resolution label map, thus supervising the local consistency. The Macro (Micro) discriminator yields ``fake'' if semantic (local) inconsistency is observed, otherwise it gives ``real''. \textbf{Bottom}: qualitative results of using Macro $D$, Micro $D$ and MMAN, respectively. We observe that Macro $D$ and Micro $D$ correct semantic inconsistency (\textcolor{green}{green} dashed circle) and local inconsistency (\textcolor{red}{orange} dashed circle), respectively, and that MMAN possesses the merits of both.}
\caption{MMAN has three components: a dual-output generator (\textcolor{blue}{blue} dashed box), a Macro discriminator (\textcolor{green}{green} dashed box) and a Micro discriminator (\textcolor{red}{orange} dashed box). Given an input image of size $3\times256\times256$, the generator $G$ first produces a low-resolution ($8192\times16\times16$) tensor, from which a low-resolution label map ($C\times16\times16$) and a high-resolution label map ($C\times256\times256$) are generated, where $C$ is the number of classes. Finally, for the each label map (sized $C\times16\times16$, for example), we concatenate it with an RGB image (sized $3\times16\times16$) along the 1st axis (number of channels), which is fed into the corresponding discriminator.}
\caption{Comparison of human parsing accuracy on the PPSS dataset~\cite{luo2013pedestrian}. Best performance is highlighted in \textcolor{blue}{blue}.}
\caption{Data available for acoustic modelling a) Train data b) Test data (Somali). \textcolor{red}{[trn: Considder making (b) extra 2 columns of (a), with dashes for all languages except Somali.]}}
\caption{Data available for language modelling. \textcolor{red}{[trn: Can you expand this table to list word types and word tokens and number of sentences for each of: Web text, Facebook posts, Facebook comments, Subttotal, Artificial text, Total]}}
\caption{Error rates ($\%$) of different methods on CIFAR-10, CIFAR-100, and SVHN datasets. The best records of our models are in \textbf{bold} and the best results are highlighted in {\color{red} \textbf{red}}.}
\caption{(Color online) Collective $\alpha$-relaxation times of both species as a function of reduced wave number $q^*$ at the state points $a5$($\CIRCLE$-$\tau_b^{\alpha}$, $\Circle$-$\tau_s^{\alpha}$) and $b5$ (\textcolor{antiquebrass}{$\blacktriangle$}-$\tau_b^{\alpha}$, \textcolor{antiquebrass}{$\bigtriangleup$}-$\tau_s^{\alpha}$), as predicted by the SCGLE approximation.}
\caption{\color{Gray}\textbf{a) PADME-ECFP architecture}. The Extended-Connectivity Fingerprint was used as the molecular input to the model. \textbf{b) PADME-GraphConv architecture.} Note that the graph convolutional network generating the latent molecular vector is trained together with the rest of the network, while the protein descriptor generation process is independent of the training of the network. The black dots represent omitted neurons and layers.}
\caption{\color{Gray}Dataset sizes after filtering.}
\caption{The regression performance across the datasets measured in RMSE (smaller is better), averaged across independent repetitions of CV. The mean RMSE are enclosed in square brackets; sample standard deviations are also reported. The best results in the PADME models are boldfaced, and one-sided two-sample t-tests are conducted against them. The \textcolor{blue}{blue} values are insignificantly bigger (worse) (p $>$ 0.05) than the boldfaced values, while the \textcolor{orange}{orange} ones are insignificantly smaller (better) (p $<$ 0.95) than them. The uncolored ones are significantly worse than the boldfaced values.}
\caption{The regression performance across the datasets measured in Concordance Index (larger is better), averaged across independent repetitions of CV. Similar to \cref{tab_rmse}, the mean CI are enclosed in square brackets; sample standard deviations are also reported. One-sided two-sample t-tests are conducted against the best PADME models. The \textcolor{blue}{blue} values are insignificantly smaller (worse) (p $>$ 0.05) than the boldfaced values, while the \textcolor{orange}{orange} ones are insignificantly larger (better) (p $<$ 0.95). The uncolored ones are significantly worse than the boldfaced values.}
\caption{The regression performance across the datasets measured in \(R^2\) (larger is better), averaged across independent repetitions of CV. Similar to \cref{tab_rmse}, one-sided two-sample t-tests are conducted against the best PADME models. The \textcolor{blue}{blue} values are insignificantly smaller (worse) (p $>$ 0.05) than the boldfaced values, while the \textcolor{orange}{orange} ones are insignificantly larger (better) (p $<$ 0.95). The uncolored ones are significantly worse than the boldfaced values.}
\caption{\color{Gray}Scatter plot and contour plot of predicted VS true values across all datasets. The panels \textbf{a, b, c} and \textbf{d} correspond to \textbf{Davis, Metz, Kiba} and \textbf{ToxCast} datasets, respectively. The axes in the two plots of the same panel are the same, and both plots are generated from the same data. The diagonal lines in the scatter plots are the reference lines where \emph{predicted = true value}.}
\caption{\color{Gray}Plots for Davis dataset predicted value VS true value. Panel (a) corresponds to the true active values, while panel (b) corresponds to true inactive values. Similar to figure \ref{fig3}, all plots in the same panel are plotted from the same data.}
\caption{\color{Gray}Similar to Figure \ref{fig5}, plots for the ToxCast dataset. Panel (b) uses a different hexagon plot from (a), because that form of hexagon plot on panel (b) does not display properly.}
\caption{\label{fig: lambda}Upper bounds $C$ on the maximum of $x$ over the global attractor of the Lorenz equations at the standard parameters. The bounds are optima of the inner minimization in \eqref{eq: opt sos} for various $\lambda$ and $V$ of polynomial degrees 2 (\dottedrule), 4 (\dashdottedrule), 6 (\longdashedrule), and 8 (\solidrule).}
\caption{\label{fig: boundaries}Attracting sets $\Omega_V^C$ that give our best upper bounds $C$ on the maximum of $x$ over the global attractor of the Lorenz equations, for Lyapunov functions $V$ of degree 2 (\dottedrule) and 4 (\dashdottedrule). Each pair $(V,C)$ solves the inner minimization in \eqref{eq: opt sos} with $\lambda=0.5659$ and $\lambda=0.3743$, respectively. The plotted curves are boundaries of the projections of $\Omega_V^C$ onto the $xz$-plane. Also shown are numerically integrated trajectories (${\color{mygray}\solidrule}$) starting along each half of the origin's unstable manifold.}
\caption{\label{fig: multi}Phase portrait of \eqref{eq: multi} showing the attracting (\solidrule) and repelling (\dashedrule) limit cycles and the repelling equilibrium ($\circ$).}
\caption{\label{fig: lambda X}Upper bounds $C$ on the maximum of $x^2$ over the minimal set $\cA_X$ that attracts all subsets of the ball $X$ of radius 4/5 for the system \eqref{eq: multi}. The bounds are optima of the inner minimization in \eqref{eq: opt X sos} for various $\lambda$ and for $V$ of polynomial degrees 2 (\dottedrule), 4 (\dashdottedrule), and 6 (\longdashedrule).}
\caption{(a)~Speed histograms; peaks correspond to escalators, stairs, and walking. (b)~the histogram for one data set with escalator rides/walking. (c--d) the histogram for roll and yaw. (e)~the paths for \textcolor{white!60!black}{ground-truth} (\ref{addplot:cs01-comparison0}), \textcolor{mycolor1}{ARKit} (\ref{addplot:cs01-comparison1}), \textcolor{mycolor2}{ARCore} (\ref{addplot:cs01-comparison2}), \textcolor{mycolor3}{Tango/Raw} (\ref{addplot:cs01-comparison3}), \textcolor{mycolor4}{Tango/Area learning} (\ref{addplot:cs01-comparison4}), \textcolor{mycolor5}{ROVIO} (\ref{addplot:cs01-comparison5}), and \textcolor{mycolor6}{PIVO}~(\ref{addplot:cs01-comparison6}).}
\caption{Example paths showing \textcolor{white!60!black}{ground-truth} (\ref{addplot:outdoor0}), \textcolor{mycolor1}{ARKit} (\ref{addplot:outdoor1}), \textcolor{mycolor2}{ARCore} (\ref{addplot:outdoor2}), \textcolor{mycolor3}{Tango/Raw} (\ref{addplot:outdoor3}), and \textcolor{mycolor4}{Tango/Area learning} (\ref{addplot:outdoor4}) that stopped prematurely in (a). Map data {\copyright}\OpenStreetMap. The ground-truth fix points were marked on an architectural drawing. ROVIO and PIVO diverge and are not shown.}
\caption{Cumulative distributions of position error: \textcolor{mycolor1}{ARKit}~(\ref{addplot:position-cdf0}), \textcolor{mycolor2}{ARCore}~(\ref{addplot:position-cdf1}), \textcolor{mycolor3}{Tango/Raw} (\ref{addplot:position-cdf2}), \textcolor{mycolor4}{Tango/Area learning} (\ref{addplot:position-cdf3}), \textcolor{mycolor5}{ROVIO} (\ref{addplot:position-cdf4}), and \textcolor{mycolor6}{PIVO}~(\ref{addplot:position-cdf5}).}
\caption{Quantitative comparison with state-of-the-art methods on six benchmark datasets. Each cell (from up to down) contains max F-measure (higher better), and MAE (lower better). The top two results are highlighted in {\color{red} \textbf{red}} and \textcolor[rgb]{0.13,0.55,0.13}{\textbf{green}} respectively. ``RA'' denotes the proposed reverse attention, and ``MK'' is MSRA-10K~\cite{Cheng43}, the other abbreviations are the initials of each dataset metioned in the paper. Note that the number of images listed here are including the augmented ones.}
\caption{This figure lists several images with generated captions relying on various parts of RNN-2DS's states. The accessible part is marked with \textcolor{blue}{blue} color in each case.}
\caption{This figure shows some qualitative samples of captions generated by different decoders, where words in \textcolor{red}{red} indicate they are inconsistent with the image.}
\caption{This figure lists some images with generated captions before and after some word-associated channel being deactivated. The word that associates with the deactivated channel is marked in \textcolor{red}{red}.}
\caption{The results, in terms of different metrics, obtained using RNN-2DS (\textcolor{green}{green}) and LSTM-1DS (\textcolor{red}{red}) on the MSCOCO offline test set with similar parameter sizes. Specifically, RNN-2DS of sizes 10.57M, 13.48M and 21.95M have compared to LSTM-1DS of sizes 10.65M, 13.52M and 22.14M.}
\caption{The traditional ToF processing pipeline ({\color{ForestGreen}green}, libfreenect2~\cite{libfreenect2} as an example) and the proposed framework ({\color{Red}red}). The lower left panel shows artifacts generated by MPI and motion, that show up respectively as deformation close to corners and spikes or missing data close to the boundaries of moving objects in the traditional flowchart ({\color{ForestGreen}green}). These artifacts are greatly reduced by the proposed framework ({\color{Red}red}), which is based on the introduction of the motion and multi-reflection modules depicted in the right part.}
\caption{Depth error for scenes from the FLAT dataset, corrupted by shot noise and MPI and reconstructed by DToF~\cite{Jar17}, Phasor Imaging~\cite{Gup15}, LF2~\cite{libfreenect2} and MRM, in cm. Errors are computed only in the unanbiguous reconstruction range of Phasor Imaging, [1.5m, 5m]; no mask is used to remove unreliable pixels. The {\color{Blue}blue} boxes in the first row show the receptive field for DToF and MRM. %\QG{{\color{Blue}Blue} boxes in the second row show the size of the receptive field for DToF and our method. Our method happens to remove long range MPI shown on the front shield of the car, despite relatively small receptive field.} }
\caption{Latent variable recovery results for the latent causal graph shown in {\color{orange}orange }and triangle causal graph shown in {\color{blue}blue}, where $\text{card}(X)=m,\text{card}(Y)=n,\text{card}(Z)=k$. For $n=m=20$, entropy of the latent variable recovered by \emph{LatentSearch} that makes $X$ and $Y$ conditionally independent given the latent is shown against the entropy of the true latent. In $(a)$ and $(b)$, points above the red line are samples for which there is no latent that makes $X,Y$ conditionally independent. As observed, there is a region of low-entropy latent variable regime for which \emph{LatentSearch} can be used to distinguish between the two causal graphs with a properly chosen threshold, e.g., if the entropy of the true latent is less than 3 bits for $|X|=|Y|=20$. }
\caption{ % The experimental effective anisotropy constant $\left| K_{\textrm{eff}} \right|$ of the \fecofivep{} system at 3~K % (the sign of $K_{\textrm{eff}}$ is not considered), % together with the magnetocrystalline anisotropy energy MAE values as calculated with the FPLO14 code. % In calculations the supercell method for modeling of chemical disorder and the PBE functional were used. % The relativistic effects were treated in a full 4-component formalism (including spin-orbit coupling). % (for $x$ equal to 0.2 and 0.8 several inequivalent supercells are considered). % For comparison the value of $K_1$ measured by Lamichhane et al. for \fefivep{}.~\cite{lamichhane_study_2016} % } \label{fig:fecopb2_Keff} \end{figure} % Figure~\ref{fig:fecopb2_Keff} presents the MAE($x$) dependence for the \fecofivep{} system as calculated with use of the supercell method. % The MAE calculations based on the supercell method proved to be one of the most accurate method for evaluation the MAE.~\cite{dane_density_2015} % However, our calculations were limited by computational challenges of the supercell method. % Thus, in practice we were able to consider only a relatively small number of configurations, see Sec.~\ref{subsec:comp_details}. % The scattering of individual data points for $x = 0.2$ and $x = 0.8$ is in a similar range as observed by D\"ane~\textit{et al.}~\cite{dane_density_2015} or Steiner~\textit{et al.}~\cite{steiner_calculation_2016} and shows that an averaging for several configurations is needed for accurate results. % In Fig.~\ref{fig:fecopb2_Keff} the regions of positive and negative MAE (of perpendicular and in-plane anisotropy) are separated at Co concentration $x\simeq0.5$. % The calculated MAE is equal 0.52~MJ\,m$^{-3}$ for \fefivep{} and -0.51~MJ\,m$^{-3}$ for \cofivep{}. % Whereas, the anisotropy value close to zero, observed for $x\simeq0.5$, indicates a good soft magnetic material. % Figure~\ref{fig:fecopb2_Keff} presents also the low temperature measurements of the effective anisotropy constant $\left|{K}_{\mathrm{eff}} \right|$ carried out at 3~K for several \fecofivep{} compositions within the boundaries of $0.0 \leq x \leq 0.7$. % The value of $\left| K_{\textrm{eff}} \right|$ is the highest (0.94~MJ\,m$^{-3}$) for \fefivep{} and the the lowest for a Co concentration $x\sim0.6$. % $\left|K_{\textrm{eff}} \right|$ measured for \fefivep{} is significantly larger than the $K_1$~=~0.5~MJ\,m$^{-3}$ measured at 2~K for the single crystal.~\cite{lamichhane_study_2016} % The decrease of $\left| K_{\textrm{eff}} \right|$ with $x$ is in agreement with the previous measurements for (Fe$_{0.8}$Co$_{0.2}$)$_5$PB$_2$ suggesting that 20\% Co substitution reduces the anisotropy field.~\cite{mcguire_magnetic_2015} % Previously we also showed the corresponding $\left|{K}_{\mathrm{eff}} \right|$ results for the Fe$_5$Si$_{1-x}$P$_{x}$B$_2$ system.~\cite{hedlund_magnetic_2017} % The presented values of $\left| K_{\textrm{eff}} \right|$ for \fefivep{} were $\sim$0.9~MJ\,m$^{-3}$ at 10~K and $\sim$0.65~MJ\,m$^{-3}$ at 300~K.~\cite{hedlund_magnetic_2017} % Notice that LAS is unable to determine the sign of $\left|K_{\textrm{eff}} \right|$ and thus the negative values of MAE predicted for $x \gtrsim 0.6$ cannot be confirmed by this method. % Other methods, such as magnetometry measurements in different directions for single crystals or torque magnetometry would be preferable. % Here, single crystals were not available, and up to 10 wt\% of impurities were present in the samples. % Therefore, given the limitation in the model and the starting material the results presented from these should be seen as semi-quantitative. % Taking into account the limitations of the LAS and the supercell method, the differences between theoretical and measured MAE($x$) results are acceptable. % We conclude, that Co alloying of \fefivep{} is not a good strategy to increase the MAE of this system. %-----------M vs H----------figure and description \begin{figure} \centering \includegraphics[trim = 25 30 25 30,clip,width=1.0\columnwidth]{fe5pb2_m_vs_h.ps} \caption{ % Magnetization ($M$) as a function of applied field ($H$) measured for \fefivep{} at 3~K. % The inset shows a normalized magnetization ($M/M_\mathrm{S}$) as a function of 1/$H^2$. % } \label{fig:fecopb2_m_vs_h} \end{figure} % A typical magnetization ($M$) \textit{versus} applied field ($H$) curve measured at 3~K is shown in Fig.~\ref{fig:fecopb2_m_vs_h}. % The inset of Fig.~\ref{fig:fecopb2_m_vs_h} presents a plot of $M/M_\mathrm{S}$ \textit{versus} 1/$H^2$ as used to determine the $\left| {K}_{\mathrm{eff}} \right|$ within the LAS method. % More details on the implementation of the LAS method can be found in Sec.~\ref{subsec:exp_details}. \subsection{Doping Fe$_5$PB$_2$ with 5$\boldsymbol{d}$ Elements}\label{subsec:5d} % One of the methods of tailoring the MAE is doping with 5$d$ elements.~\cite{edstrom_magnetic_2015,khan_site_2018} % Previously, we have confirmed that the 5$d$ elements can significantly affect the MAE due to a large spin-orbit coupling.~\cite{edstrom_magnetic_2015} % From the Fe$_5$Si$_{1-x}$P$_{x}$B$_2$ and \fecofivep{} systems, the highest MAE is found in the \fefivep{} phase.~\cite{hedlund_magnetic_2017} % Thus, it is considered as the parental compound for a further MAE engineering. % The MAE of (Fe$_{0.95}$X$_{0.05}$)$_5$PB$_2$ compounds (X~=~5$d$ elements) is calculated using the supercell method. % % top left bottom right % \begin{figure} \centering \includegraphics[trim = 30 75 250 110,clip,height=\columnwidth,angle=270]{fe5pb2.MAE_vs_5d.ps} \caption{ MAE for various 5$d$ elements X in (Fe$_{0.95}$X$_{0.05}$)$_5$PB$_2$ as calculated with supercell method. % Calculations were done with the FPLO14 code using the PBE functional and treating the relativistic effects in a full 4-component formalism (including spin-orbit coupling). % The dashed line indicates the MAE of \fefivep{} (0.52~MJ\,m$^{-3}$) for comparison. } \label{fig:fepb2_5d_MAE} \end{figure} % The results are shown in Fig.~\ref{fig:fepb2_5d_MAE}, with the 5$d$ element marked on the $x$ axis and dashed line indicating the MAE of undoped Fe$_5$PB$_2$. % The 5$d$ doping has sometimes beneficial and sometimes adverse effect on MAE.~\cite{ayaz_khan_magnetocrystalline_2016, edstrom_magnetocrystalline_2017,edstrom_theoretical_2016} % Significant increase of MAE is observed for W or Re doping, similar like in the case of (Fe$_{1-x}$Co$_{x}$)$_2$B alloys investigated experimentally in our previous work~\cite{edstrom_magnetic_2015}. % The MAE grows from 0.52~MJ\,m$^{-3}$ for Fe$_5$PB$_2$ to about 1.1~MJ\,m$^{-3}$ for the compositions with W or Re, with 5\% Fe substitution. % Previously we have shown, that the increase in MAE observed for W and Re dopants is mainly due to the strong spin-orbit coupling of the 5$d$ atoms, however other variations in electronic structure also affect the MAE.~\cite{edstrom_magnetic_2015} % % % top left bottom right \begin{figure} \centering \includegraphics[trim = 30 55 255 110,clip,height=\columnwidth,angle=270]{fe5pb2.mm_vs_5d.ps} \caption{ Spin ($m_\mathrm{s}$) and orbital ($m_\mathrm{l}$) magnetic moments of 5$d$ transition metal impurities X in (Fe$_{0.95}$X$_{0.05}$)$_5$PB$_2$ as calculated for spin quantization axis along the $c$-axis. % Supercell calculations were done with the FPLO14 code using the PBE functional and treating the relativistic effects in a full 4-component formalism (including spin-orbit coupling). } \label{fig:fepb2_5d_mm} \end{figure} % Although in our calculations the 5$d$ elements are initially considered as non-magnetic, the dopants undergo spin polarization in a ferromagnetic medium and contribute to the total magnetic moment of the system. % The calculated spin and orbital magnetic moments on 5$d$ impurity show clear trend along the increasing atomic number of 5$d$ element, see Fig.~\ref{fig:fepb2_5d_mm}. % The spin magnetic moment of 5$d$ impurities are antiparallel to the Fe moments in the early 5$d$ series, while they are parallel in the late 5$d$ series. % Corresponding trends for 5$d$ atoms in magnetic 3$d$ hosts have been found previously computationally~\cite{akai_nuclear_1988, dederichs_ab-initio_1991} and experimentally.~\cite{wienke_determination_1991} \section{Summary and Conclusions}\label{sec:conclulsions} % %----------------- fe5pb2 ---------------------- % Our considerations began with a detailed theoretical analysis of the \fefivep{} compound. % The Fe 3$d$ orbitals are dominant in the valence band and responsible for the formation of large magnetic moments. % For the \fefivep{} the fully relativistic band structure in the vicinity of Fermi level was considered to better understand the origin of the high value of magnetocrystalline anisotropy energy (MAE). % The calculated Fermi surface requires experimental confirmation. % %-----------------------FSM------------------------- % The results of fully relativistic fixed spin moment calculations suggested that reduction of the magnetic moment of \fefivep{} should induce about fourfold increase of the MAE. % For practical realization of magnetic moment reduction it is suggested to alloy Fe with a non-magnetic element Ru or Os from the Fe group, or to partially replace Fe with two elements at once, Cr and Ni, for example, keeping constant number of valence electrons. %---------------------(Fe1-xCox)5PB2------------ % Three critical parameters for technological applications: saturation magnetization ($M_\mathrm{S}$), Curie temperature ($T_\mathrm{C}$), and MAE were calculated for the whole concentration range between \fefivep{} and \cofivep{}. % The calculated $M_\mathrm{S}$ and $T_\mathrm{C}$ decreased with Co concentration and for the terminal composition \cofivep{} a weakly ordered magnetic ground state was predicted. % The calculated $M$($x$) and $T_\mathrm{C}$($x$) were in decent agreement with the measurements, although the ferromagnetic ground state of \cofivep{} is questionable. % The Co doping in \fecofivep{} system gives the possibility of tuning the $T_\mathrm{C}$ in a range from about six hundred kelvins to almost down to zero. % The calculated MAE was positive for \fefivep{}, negative for \cofivep{}, and went through zero around 50\% Co concentration. % %--------------------exp. K_eff----------------------- % This picture of MAE($x$) behavior was in overall agreement with the experimental study of the effective anisotropy constant $\left|K_{\textrm{eff}} \right|$ for the \fecofivep{} alloys. % The measurements showed the highest $\left| K_{\textrm{eff}} \right|$ value for stoichiometric \fefivep{} which decreased with Co doping. % We concluded then that Co alloying is not a good strategy to increase the MAE of \fefivep{} alloy. % The measured $\left|K_{\textrm{eff}} \right|$ of about 0.94~MJ\,m$^{-3}$ at 3~K was, however, the highest value obtained so far for \fefivep{}, giving a hope for potential application of its other alloys. % It was also calculated how the 5\% doping of Fe with 5$d$ elements affects the MAE of the \fefivep{}. % It was shown that \fefivep{} doping with W or Re results in significant increase of the magnetocrystalline anisotropy energy. \begin{acknowledgments} % MW and JR acknowledge the financial support from the Foundation of Polish Science grant HOMING. % The HOMING programme is co-financed by the European Union under the European Regional Development Fund. % JC and MS acknowledge the financial support from the Swedish Research Council. % DH, PS, KG acknowledge Swedish Foundation for Strategic Research for financial support. % Part of the computations were performed on resources provided by the Poznań Supercomputing and Networking Center (PSNC). % We thank Bartosz Wasilewski for help with language editing and Dr Jakub Kaczkowski for reading the manuscript and helpful discussion. % \end{acknowledgments} \end{sloppypar} \bibliography{feco5pb2_MAE} % feco5pb2_MAE.bib is the name of our database \end{document}}
\caption{Analysis on case 5a: subplot (a) - drag coefficient ({\color{red}\rule{0.5cm}{2pt}}) superimposed on the space-time diagram. Subplot (b) variation of $u'_z(0,0,z,t)$ in z direction for $t=1000\ $ (LS {\color{magenta}\rule{0.5cm}{2pt}}) and $t=2500\ $ (US2 {\color{black}\rule{0.5cm}{2pt}}). \label{Case_5a}}
\caption{Time signal of $u'_z(0,0,z,t)$ in laminar state recorded for $z=40d$ (a {\color{matlab5}\rule{0.5cm}{2pt}}) and $z=80d$ (b {\color{matlab4}\rule{0.5cm}{2pt}}). Time signal of $u'_z(0,0,z,t)$ in US2 recorded for $z=40d$ (d {\color{matlab1}\rule{0.5cm}{2pt}}) and $z=80d$ (e {\color{matlab2}\rule{0.5cm}{2pt}}). The position of plots in this figure are marked with the corresponding colour in figure (\ref{Case_5a}a).\label{Case_5abis}}
\caption{Correlation between $(L_R^*)^2$ and $A_R^*$. The grayscale indicates the value of $St_A$ and the dashed line (\protect\dashedrule) represents a best-fit of Eq. \ref{eq:AR_scaling_2} on available data.}
\caption{Mass balance along the mean recirculation region. Symbols: \protect\graysquare $\left|\epsilon_R^*\right|$; \protect\dashedrule $\left|\epsilon_R^*\right|= 0.025$; \protect\whitecircle $\dot{m}_R^*$. Empty symbols indicate points for which $\left|\epsilon_R^*\right|< \dot{m}_R^*$.} \label{fig:mass_error} \end{figure} As shown in Figure \ref{fig:mass_error}, the recirculation region verifies Eq. \ref{eq:massBalance} approximately within an error $\epsilon_R^* \approx \pm$ \num{2.5e-2} for all tested values of $St_A$, which seems acceptable, at least in comparison to errors reported by \citet{stella2017}. A large fraction of $\epsilon_R^*$ seems to be due to laser reflections at the wall, that produce spurious vectors in regions of the velocity field that surround the extrema of the mean separation line. It is clear from Figure \ref{fig:vortex} that the amount of mass that rotates with the vortex should alone assure continuity. Eq. \ref{eq:massBalance} can then be recasted as: \begin{equation} \dot{m}_R^{* \, IN} + \dot{m}_R^{* \, OUT} = 0, \label{eq:massBalance_2} \end{equation} where the superscripts $^{IN}$ and $^{OUT}$ indicate mass injection and extraction, respectively. According to Eq. \ref{eq:massBalance_2}, then, the estimation of the backflow $\dot{m}_R^*$ requires to condition the velocity integral of Eq. \ref{eq:massBalance} to the sign of $V$, which gives: \begin{equation} \dot{m}_R^* = \dot{m}_R^{* \, IN} = - \frac{\int_{S_R}V^{-}cos\left(\phi\right) ds}{U_\infty h} = - \dot{m}_R^{* \, OUT} = \frac{\int_{S_R}V^{+}cos\left(\phi\right) ds}{U_\infty h}. \label{eq:backflow} \end{equation} In this expression, $V^{+}$ represents positive velocity values and $V^{-}$ negative ones, respectively. In itself, continuity does not put any constraint on the form of the integrands of Eq. \ref{eq:backflow}. Anyway, results reported by \citet{stella2017} for the unperturbed flow show that $V^{+}$ are rather concentrated over $x/h < L_R^*/2$, while $V^{-}$ are found on $x/h > L_R^*/2$. This is also well verified in the present experiment, at least if $L_R^*$ is not too small, as suggested by shape of the streamlines shown in Figure \ref{fig:vortex}. In addition, phenomenological arguments supporting such an approximately odd distribution of $V$ along the mean separation line can be found in the work of \citet{chapman1958}: these researchers highlighted that the amount of mass \textit{scavenged} into the recirculation region at reattachment should balance the flux of mass re-entrained into the shear layer at separation. The evolution of $\dot{m}_R^*$ with $St_A$, computed using Eq. \ref{eq:backflow}, is presented in Figure \ref{fig:Freq_resp} and Figure \ref{fig:mass_error}. Interestingly, Figure \ref{fig:Freq_resp} shows that $\dot{m}_R^*$ is approximately inversely correlated to the frequency response of the jet, even if the minimum value of $\dot{m}_R^*$ seems to be attained for higher values of $St_A$ than those for which $U_{JP}^*$ reaches its maximum. In any case, it appears that, in first approximation, the higher is $U_{JP}^*$, the more effective is the jet at hindering the backflow. It is important to remark that even if the value of $\epsilon_R^*$ is quite homogeneous across the spanned $St_A$ domain, its relative impact on our investigation increases as $\dot{m}_R^* \rightarrow 0$. In particular, it is $\epsilon_R^* > \dot{m}_R^*$ on $St_A \in (0.12,0.22)$. For sake of safety, then, datapoints within this frequency subrange (highligted by empty symbols in Figure \ref{fig:mass_error}) will be discarded in the remainder of the paper. \subsection{Backflow and vortex circulation}\label{sec:backflowCirc} Now that values of $\dot{m}_R^*$ are available, we can start the discussion of the vortex model by investigating whether $\dot{m}_R^*$ can be related to $\Gamma_V^*$. Let us begin by making the assumption that the total circulation of the flow can be modelled as: \begin{equation} \Gamma^* = \Gamma_0^* + \Gamma_V^*, \label{eq:gamma_Tot} \end{equation} where $\Gamma_0^*$ is the circulation of a hypothetical flow with no separation. In principle, $\Gamma_0^*$ should only depend on geometry and free flow velocity. In the reference system of Figure \ref{fig:rampaGDR}, it is both $\Gamma_0^* < 0$ and $\Gamma_V^*<0$. Eq. \ref{eq:gamma_Tot} implies that all variations of $\Gamma^*$ measured in the experiment should be attribuable to variations of the properties of the vortex. In this respect, it does not seem unreasonable to put: \begin{equation} \Gamma_V^*(St_A) \sim \Gamma_{jet}^*(St_A), \label{eq:gammaV_gammaJ} \end{equation} as the size of the vortex is driven by the action of the synthetic jets (see Figure \ref{fig:LR_StA}). Mind that the correlation postulated in Eq. \ref{eq:gammaV_gammaJ} appears to be a negative one. Indeed, since the velocity scale of the free flow remains $U_\infty$ (see also Sec. \ref{sec:wallNormPrexGrad} for what concerns the recirculation region), $\Gamma_V^*$ should decrease as $L_R^*$ shrinks and hence, according to the analysis proposed by \citet{berk2017}, as $\Gamma_{jet}^*(St_A)$ increases. The same work by \citet{berk2017} suggests that the mean amount of circulation injected in the flow by the controller is given by: \begin{equation} \Gamma_{jet}^* \sim \frac{1}{T U_\infty^2}\int_0^T \frac{1}{2}U_J^2(t,St_A)\mathrm{d}\tau \sim (U_{JP}^*)^2 \sim \left(\dot{m}_R^*\right)^2, \label{eq:gammaJet} \end{equation} in which $\tau$ indicates time, $T$ is the actuation period and the dependency on $St_A$ was omitted for simplicity. According to Figure \ref{fig:Freq_resp}, the proportionality with respect to $\dot{m}_R^*$ should hold at least on $St_A > 0.1$. By plugging Eq. \ref{eq:gammaJet} into Eq. \ref{eq:gamma_Tot} one obtains: \begin{equation} \Gamma^* - \Gamma_0^* = \Gamma_V^* \sim \left(\dot{m}_R^*\right)^2, \label{eq:gamma_mR} \end{equation} In a bidimensional flow, $\Gamma^*$ is classically computed as: \begin{equation} \Gamma^* = \frac{1}{U_\infty h}\oint_{S_V} \boldsymbol{U}\cdot\boldsymbol{t} \, ds, \label{eq:gammaDef} \end{equation} where $\boldsymbol{t}$ is the local tangent to the closed contour $S_V$, which is presented in Figure \ref{fig:vortex}. As for what concerns $\Gamma_0^*$, it can be estimated by fitting Eq. \ref{eq:gamma_mR} onto available values of $\Gamma^*$ and $\dot{m}_R^*$, which gives $\Gamma_0^* \approx -5.21$. Figure \ref{fig:GammaV_mR} shows that, in spite of some scatter, computed values of $\Gamma_V^*$ verify Eq. \ref{eq:gamma_mR} relatively well. It seems then safe to consider that $\dot{m}_R^*$ is related to $\Gamma_V^*$, i.e. that the backflow can be assimilated to the amount of mass put in rotation by the vortex. Since the vortex dominates the separated flow, the link with $\Gamma_V^*$ gives strong support to the idea that $\dot{m}_R^*$ is a key quantity in the functionning of separating/reattaching flows. \begin{figure} \centering \includegraphics[width=\linewidth]{./fig10} \caption{Evolution of $\Gamma^* - \Gamma_0^*$ with respect to $\left(\dot{m}_R^*\right)^2$. The grayscale indicates the value of $St_A$ and the dashed line (\protect\dashedrule) represents a best-fit of Eq. \ref{eq:gamma_mR} on available data.} \label{fig:GammaV_mR} \end{figure} \subsection{Backflow and vortex size}\label{sec:LR_mR} The second vortex parameter that has to be compared to $\dot{m}_R^*$ is its characteristic scale $L_R^*$. Not surprisingly, Figure \ref{fig:LR_mR} shows that $\dot{m}_R^*$ and $L_R^*$ are directly correlated, as could be expected by the fact that $L_R^*$ scales the exchange surface of the mean recirculation region (see Eq. \ref{eq:backflow} and discussion). The trend of $L_R^*$ with respect to $\dot{m}_R^*$ appears to be well approximated by a linear model in the form: \begin{equation} L_R^* = k_{m,L}\dot{m}_R^* + L_{R,0}^*, \label{eq:LR_mR} \end{equation} where $k_{m,L}$ and $L_{R,0}^*$ were estimated at approximately \num{60} and \num{0.8}, respectively, by fitting Eq. \ref{eq:LR_mR} on available data. The non-null value of $L_{R,0}^*$ is somehow surprising, because it seems to imply that the recirculation region does not disappear for $\dot{m}_R^* \rightarrow 0$. In this paper we do not tackle this problem directly, but some arguments supporting $L_{R,0}^*>0$ are discussed in Appendix. Regardless to the value of $L_{R,0}^*$, anyway, Eq. \ref{eq:LR_mR} is in fundamental agreement with findings reported by \citet{berk2017}. Indeed, comparison of Figure \ref{fig:LR_mR} with Figure \ref{fig:LR_StA} proves that $\dot{m}_R^*$ is a more appropriate predictor of the evolution of controlled separating/reattaching flows than $St_A$ \cite{berk2017}, providing a sensibly simpler, monotonical description of the behaviour of the characteristic length scale $L_R^*$. It is stressed that the linear trend presented in Figure \ref{fig:LR_mR} was obtained without making any hypothesis on the characteristics of the actuator. In fact, observations reported by \citet{stella2017} on the scaling of $L_R^*$ in a unperturbed flow suggest that a linear relationship between $\dot{m}_R^*$ and $L_R^*$, similar to Eq. \ref{eq:LR_mR}, might be a general property of separating/reattaching flows assimilable to the one under study. \begin{figure} \centering \includegraphics[width=\linewidth]{./fig11} \caption{Evolution of $L_R^*$ with respect to $\dot{m}_R^*$. The grayscale indicates the value of $St_A$ and the dashed line (\protect\dashedrule) represents a linear fit as in Eq. \ref{eq:LR_mR}.} \label{fig:LR_mR} \end{figure} \section{An estimator for $\boldsymbol{\dot{m}_R^*}$}\label{sec:backflowObs} By exploiting the vortex model, previous sections contributed at affirming the key role of mass entrainment in separating/reattaching flows. In this regard, two findings seem particularly promising in the perspective of separation control. In the first place, it appears that $L_R^*$ (i.e. the \textit{state} of the flow, at least for what concerns topological analysis) could be reconstructed simply from the backflow $\dot{m}_R^*$. In the second place, the inverse correlation between $\dot{m}_R^*$ and the characteristic actuation velocity $U_{JP}^*$ might hold promise of entrainment-based, closed-loop control systems, in which $\dot{m}_R^*$ feedback might be used to regulate the intensity of the command given to attain a target flow state. In the light of these elements, in principle $\dot{m}_R^*$ stands out as a powerful control variable, that might provide both information to reconstruct the state of the flow and input for actuation. Unfortunately, in practice measuring $\dot{m}_R^*$ is impossible in most real-life applications, because the large velocity fields that are necessary to its computation are usually not available (see Sec. \ref{sec:backflowComp}). In this respect, we argue that the vortex model might provide a model-based definition of simply deployable observers for $\dot{m}_R^*$, and hence $L_R^*$. \subsection{Relating $\boldsymbol{\dot{m}_R^*}$ to the pressure field}\label{sec:backFlow2Prex} Practical problems in sensing industrial flows make it suitable to base the estimation of $\dot{m}_R^*$ on simply accessible information. The first candidate quantity that comes to mind is wall-pressure, which in most applications can be directly measured with relatively unexpensive, available on-the-shelf, flush-mounted pressure taps. The mean wall-pressure distribution typical of separating/reattaching flows assimilable to the one under study is well characterised for the baseline flow (for example, see \cite{roshko1965} and related, subsequent literature) and already documented in the case of flows controlled with synthetic jets \cite{guilmineau2014}. Anyway, a quantitative link between mean wall-pressure (and the pressure field more in general) and $\dot{m}_R^*$ is not self evident. It is then convenient to begin our discussion by explicitely investigating if such link exists. In this regard, the vortex model can help our reasoning, as follows. Since the vortex dominates the mean flow, it does not seem unreasonable to relate the vertical pressure force acting on the mean separation line to the circulation $\Gamma_V$. By invoking the Joukowski theorem, this relation can be expressed as: \begin{equation} \rho U_\infty\Gamma_V \approx \int_{S_R}\left(P(s)-P_\infty\right) ds \approx \left(P_V-P_\infty\right)S_R \sim \left(P_V-P_\infty\right)L_R, \label{eq:Jukou1} \end{equation} in which $P(s)$ is the pressure distribution along the mean separation line, $P_V$ is its average value and $P_\infty$ is the mean static pressure in the free stream above the descending ramp. It suits our purposes to consider that $P_V \approx P(x_V,y_V)$, where $x_V$ is the streamwise position of the vortex center (see Figure \ref{fig:vortex}). At least for what concerns the baseline flow, this hypothesis is supported by the odd form of the pressure gradient reported by \citet{stella2017}. By normalising all quantities in Eq. \ref{eq:Jukou1}, one naturally obtains: \begin{equation} \Gamma_V^* \sim C_{p,V}L_R^*. \label{eq:Jukou3} \end{equation} In this expression, $C_{p,V}$ can be interpreted as the characteristic pressure coefficient of the center of the vortex, computed as: \begin{equation} C_{p,V} = \frac{P_V - P_\infty}{1/2 \rho U_\infty^2}. \label{eq:Cp} \end{equation} Starting from the vortex model, in previous sections we have already proposed dependencies on $\dot{m}_R^*$ for both $\Gamma_V^*$ (Eq. \ref{eq:gamma_mR}) and $L_R^*$ (Eq. \ref{eq:LR_mR}). By plugging these expressions into Eq. \ref{eq:Jukou3}, one simply obtains: \begin{equation} C_{p,V} \sim \frac{\Gamma_V^*}{L_R^*} \sim \dot{m}_R^*, \label{eq:vortexPrex} \end{equation} which should hold at least if $L_R^*$ is not too small with respect to $L_{R,0}^*$. Eq. \ref{eq:vortexPrex} provides a first connection between the pressure field and the backflow, even if $C_{p,V}$ is generally not accessible without deeply perturbing the separated flow. We then need to go one step further in our reasoning, and relate $C_{p,V}$ to a wall-pressure value. \subsection{The wall-normal pressure gradient within the recirculation region}\label{sec:wallNormPrexGrad} In order to introduce mean wall-pressures into Eq. \ref{eq:vortexPrex}, it is useful to look into the effects of the external forcing on the spanwise position of the center of the vortex $x_V^*$. Figure \ref{fig:xV_Sth} shows that, with the exception of the baseline flow and of few controlled flows which are assimilable to it (e.g. $St_A \approx 1$), $x_V^*$ is relatively stable and similar to $x_{low}^* = 1/tan\left(\alpha\right)$, which is the position of the lower edge of the ramp. This suggests that $P_b$, that is the wall-pressure at the base of the ramp (see Figure \ref{fig:vortex}), might be related to the pressure field induced by the vortex, in particular at its center. For simplicity, let us put $x_V^* \approx x_{low}^*$. Then, it will be: \begin{equation} C_{p,b} \approx C_{p,V} + \int_{-1-y_V^*}^{-1} \frac{\partial C_p}{\partial y^*} \mathrm{d}y^*. \label{eq:Cpb_1} \end{equation} The non-dimensional vertical pressure gradient $\partial C_p/\partial y^*$ can be computed with the RANS equation along the wall-normal direction, as: \begin{equation} \frac{\partial C_p}{\partial y^*} \approx -\frac{2h}{U_\infty^2}\left(U\frac{\partial V}{\partial x} + V\frac{\partial V}{\partial y} - \frac{\partial \langle u^\prime v^\prime \rangle}{\partial x} - \frac{\partial \langle v^{\prime 2} \rangle}{\partial y}\right), \label{eq:dCp_dy} \end{equation} where the symbol $\langle\bullet\rangle$ indicates ensemble averaging. It is practical to estimate the order of magnitude of $\partial C_p/\partial y^*$ with some dimensional analysis. An interesting starting point is given by the relationship $b^* \sim L_R^*$ (Eq. \ref{eq:yV_LR}), which has implications on the scaling of velocities within the recirculation region. Indeed, since the backflow must remain constant through all sections of the recirculation region, the following should be verified: \begin{equation} -b^* U_R^* \approx \dot{m}_R^* \approx v_{e,R}^*\frac{L_R^*}{2}. \label{eq:mass_balance_vortex} \end{equation} In this expression, $U_R^*<0$ is a characteristic streamwise velocity scale within the recirculation region and $v_{e,R}^*$ is the mean entrainment velocity along the mean separation line. We show in Appendix that, if $L_R^* > 2.5$, $v_{e,R}^*$ is approximately independent of $St_A$ and $v_{e,R}^* = \mathcal{O}\left(10^{-2}\right)$. Of course, Eq. \ref{eq:yV_LR} implies that $U_R \sim U_\infty$ as long as $v_{e,R}^* \approx const.$ , that is, if the recirculation region is not too small, the velocity scale within the recirculation region is similar to the one of the baseline flow, regardless to the working point of the synthetic jets. If this is so, we can tentatively rely on previous works on natural separated flows to assess the order of magnitude of each term of Eq. \ref{eq:dCp_dy}. In particular, results reported by \citet{le1997}, \citet{dandois07} and \citet{stella2017} (among others) suggest that $U_R^* = \mathcal{O}\left(10^{-1}\right)$ within the recirculation region, and that the turbulent terms will tend to cancel each other out. Previous sections allow us to also assume that the characteristic horizontal and vertical length scales of the recirculation region will both depend on $L_R^*$. In addition, it seems possible to consider that $V^* \sim v_{e,R}^*$ (for this approximation, the reader is referred to Eq. \ref{eq:ve_Uinf} and \ref{eq:exchangeLenght} in appendix \ref{sec:entrRates}). With these dimensional considerations in mind, Eq. \ref{eq:dCp_dy} reduces to: \begin{equation} \frac{\partial C_p}{\partial y^*} \approx - U^*\,\frac{\partial V^*}{\partial x^*} \sim -L_R^{* -1}\mathcal{O}\left(10^{-3}\right). \label{eq:dCp_dy_2} \end{equation} Since $L_R^* = \mathcal{O}(1)$, this leads to: \begin{equation} C_{p,b} \approx C_{p,v} \sim \dot{m}_R^* \label{eq:Cpb_2} \end{equation} The evolution of $C_{p,b}$ with respect to $\dot{m}_R^*$ is presented in Figure \ref{fig:Cp_mR}, showing relatively good agreement with the linear trend predicted by Eq. \ref{eq:Cpb_2}. We stress the practical interest of this result: the base pressure $C_{p,b}$, obtainable with a single flush-mounted pressure tap, appears to be a reliable observer of $\dot{m}_R^*$. Since the main characteristic properties of the vortex model, including $L_R^*$, were shown to evolve approximately monotonically with $\dot{m}_R^*$, $C_{p,b}$ might allow to simply reconstruct many fundamental aspects of the large-scale mean topology of separating/reattaching flows, without the need of expensive and unpractical sensing systems. \begin{figure} \centering \includegraphics[width=\linewidth]{./fig12} \caption{Evolution of the streamwise position of the vortex $x_V^*$ with respect to $St_A$. All available datapoints are included, regardless to the value of $\epsilon_R^*$ (see Sec. \ref{sec:backflowComp}).} \label{fig:xV_Sth} \end{figure} \begin{figure} \centering \includegraphics[width=\linewidth]{./fig13} \caption{Evolution of $C_{p,b}$ with respect to $\dot{m}_R^*$. The grayscale indicates the value of $St_A$ and the dashed line (\protect\dashedrule) represents a linear fit.} \label{fig:Cp_mR} \end{figure} \section{Conclusions}\label{sec:concluVortex} In this study we have proposed an original model of mean separating/reattaching flows, which puts the backflow $\dot{m}_R^*$ at the core of their functionning. Our model represents the mean flow as a large spanwise vortex, covering the entire recirculation region and most of the separated shear layer. The vortex is defined by two characteristic parameters, its scale $L_R^*$ and its circulation $\Gamma_V^*$. To test the relevance of the model, we have analysed the relationship between $\dot{m}_R^*$ and the parameters of the vortex on a set of 21 different mean separated flows, each obtained by forcing a single separated ramp flow at a different actuation frequency. The external forcing, provided by a rack of synthetic jets, significantly changes the mean topology of the flow, but it does not impact its global $L_R^*$ scaling. As a first step, we have estimated $\dot{m}_R^*$ from PIV data, showing that its intensity is approximately inversely correlated to the characteristic jet velocity $U_{JP}^*$. We have also demonstated that the circulation induced by the vortex $\Gamma_V^*$ scales as $\left(\dot{m}_R^*\right)^2$: as such, $\dot{m}_R^*$ can be assimilated to the amount of mass that is entrained in rotation by the circulation of the vortex. Since in our model the vortex dominates the separated flow, this result immediately highlights $\dot{m}_R^*$ as one of the key variables driving mean separating/reattaching flows. On this basis, we have then analysed the correlation between $\dot{m}_R^*$ and $L_R^*$, showing that it can be satisfactory approximated with a linear model. This finding confirms and extends previous studies on natural and forced separating/reattaching flows, suggesting that $\dot{m}_R^*$ is a more appropriate estimator of the variation of $L_R^*$ than other flow or actuation parameters, such as the thickness of the incoming boundary layer or $St_A$. This might pave the way to a universal description of separating/reattaching flows, in particular independent of the characteristics of the actuators. As a final contribution, we have exploited the vortex model to tackle the problem of estimating $\dot{m}_R^*$ in industrial applications. By invoking the Joukovsky theorem, we have proven that $\dot{m}_R^*$ is linearly correlated to the pressure at the center of the vortex, which is itself well approximated by wall-pressure at the base of the ramp. As such, it appears that a single pressure measurement, simply accessible with a flush-mounted tap, might be sufficient to estimate $\dot{m}_R^*$. In turn, this might allow us to both reconstruct $L_R^*$ (as well as many large-scale features of the mean flow scaling with it) and close the control loop, by using $\dot{m}_R^*$ as feedback to tune $U_{JP}^*$. In the light of these findings, our future efforts will pursue two complementary objectives. In the first place, we aim at exploiting mass entrainment to develop new model-based, closed-loop separation control systems. In the second place, we would like to extend the vortex model adopted in this study, in the hope of contributing to the development of fast, inexpensive numerical tools to predict the large-scale features of separating/reattaching flows. \begin{acknowledgments} This work was supported by the CNRS Groupement De Recherche (GDR) 2502 “Flow Separation Control”, and by the French National Research Agency (ANR) through the Investissements d'Avenir program, under the Labex CAPRYSSES Project (ANR-11-LABX-0006-01). The authors wish to gratefully thank M. St\'{e}phane Loyer (PRISME, Univ. Orl\'{e}ans) for his contribution to wind tunnel measurements. \end{acknowledgments} \appendix \section*{Appendix: a note on entrainment rates}\label{sec:entrRates} Although not essential to the present discussion of the vortex model, it seems important to complete the analysis of the role of the backflow by briefly investigating the mean \textit{entrainment rate} along the mean separation line. Hereafter, this quantity will be indicated with the symbol $v_{e,R}^*$. Interest in the behaviour of $v_{e,R}^*$ is motivated by several previous findings. In particular, \citet{stella2017} shows that $v_{e,R}^*$ plays a crucial role in the development of the separated shear layer (along with the entrainment rate of external fluid from the free flow) and hence in the tuning of $L_R^*$ \cite{adamsJohnstonPart1,barros2016,berk2017}. In this respect, we believe that, through the study of $v_{e,R}^*$, our experimental database and some of the findings of previous sections can provide new, relevant insight into the interactions between the actuator and the separated flow. According to \citet{stella2017}, $v_{e,R}^*$ can be computed as: \begin{equation} v_{e,R}^* = \frac{v_{e,R}}{U_\infty}\approx \frac{\dot{m}_R^*}{S_R^*/2}. \label{eq:ve_Uinf} \end{equation} In this expression, $v_{e,R}$ is a mean entrainment velocity computed along the mean separation line, say on $x/h \in \left(L_R/2, L_R\right)$, and hence, in first approximation: \begin{equation} S_R^*/2 = \frac{1}{h}\int_{s_{V}}^{s_1} ds \sim L_R^*, \label{eq:exchangeLenght} \end{equation} where $s_V$ and $s_1$ are the values of $s$ corresponding to $x_V$ and $x_R$, respectively. Figure \ref{fig:ve_mR} presents values of $v_{e,R}^*$ yielded by Eq. \ref{eq:ve_Uinf}, for all retained values of $St_A$ (see Sec. \ref{sec:backflowComp}). Eq. \ref{eq:ve_Uinf} suggests that $v_{e,R}^*$ can be interpreted as a measure of efficiency of mass exchange through the mean separation line. Then, $v_{e,R}^*$ is plotted in function of $\dot{m}_R^*$. Results reported by \citet{stella2017} show that in a unperturbed flow $v_{e,R}^*$ is insensitive to changes in the structure of the separated shear layer, caused by a relatively wide variation of $\Rey_\theta$ in the incoming boundary layer. Quite surprisingly, a similar behaviour is also observed in Figure \ref{fig:ve_mR}, for $\dot{m}_R^* > 0.03$. On this domain, indeed, it is $v_{e,R}^* \approx v_{e,H}^*\approx 0.021$, which is in very good agreement with entrainment rates observed by \citet{stella2017} ($v_{e,R}^* \approx 0.024 \pm 0.002$). Since $\dot{m}_R^*$ is linearly correlated to $L_R^*$ (see Figure \ref{fig:LR_mR}), it appears that for large recirculation regions the efficiency of mass exchanges through the mean separation line is as insensitive to actuation as to $\Rey_\theta$. In addition, $v_{e,H}^*$ does not seem to be much affected by the value of the parameter $\delta_e/h$ either \cite{adamsJohnstonPart1}, as suggested by comparing the present experiment ($\delta_e/h \approx 0.3$) with \citet{stella2017} ($\delta_e/h \approx 0.9$). This being so, it seems likely that $v_{e,H}^*$ might rather depend on factors that were kept similar across experiments, in particular geometric characteristics such as the profile of the ramp, or its expansion ratio (approximately \num{1.1} in both cases). For $\dot{m}_R^* < 0.03$, Figure \ref{fig:ve_mR} shows that $v_{e,R}^*$ decreases linearly, approximately of a factor 10 per every decade of $\dot{m}_R^*$. Since $\dot{m}_R^*$ is also linearly correlated to $L_R^*$, Eq. \ref{eq:LR_mR} implies that the jets have an effect on $v_{e,R}^*$ only if $L_R^*<2.5$. Such small recirculation regions are attained for $St_A$ approximately within $\left(0.05,0.5\right)$, which broadly corresponds to those actuation frequencies for which $U_{JP}^* \approx U_{JP,max}^*$. According to Figure \ref{fig:LR_StA}, however, about \SI{75}{\percent} of the variation of $L_R^*$ is achieved for $St_A<0.05$ or $St_A>0.5$. In spite of the high values of $U_{JP}^*$, then, the synthetic jets become less effective at modifying $L_R^*$ when they also act on $v_{e,R}^*$. In this regard, it seems worth pointing out that the trend of $v_{e,R}^*$ shown in Figure \ref{fig:ve_mR} conceptually reminds the transfer function of a high-pass filter, in the form: \begin{equation} v_{e,R}^* \approx v_{e,H}^*\frac{\dot{m}_R^*/\dot{m}_{R,cut}^*}{1 + \dot{m}_R^*/\dot{m}_{R,cut}^*}. \label{eq:transFunct} \end{equation} In this expression, $\dot{m}_{R,cut}^*$ is a cutoff value, that can be estimated as the value of $\dot{m}_R^*$ for which $v_{e,R}^* \approx v_{e,H}^*/2$. For the present experiment, available data yield $\dot{m}_{R,cut}^* \approx 0.0123$. Let us now take a look at Eq. \ref{eq:ve_Uinf} once again. By considering $S_R^* \approx L_R^*$ and plugging in Eq. \ref{eq:LR_mR}, simple manipulations lead to: \begin{equation} v_{e,R}^* \approx 2 \frac{\dot{m}_R^*}{L_R^*} \approx \frac{2}{k_{m,L}} \,\frac{\dot{m}_R^*/\left(L_{R,0}^*/k_{m,L}\right)}{1 + \dot{m}_R^*/\left(L_{R,0}^*/k_{m,L}\right)}. \label{eq:ve_Uinf_2} \end{equation} Comparing this latter expression with Eq. \ref{eq:transFunct} gives $\dot{m}_{R,cut}^* = L_{R,0}^*/k_{m,L} \approx 0.0133$ and $v_{e,H}^* = 2/k_{m,L} \approx 0.033$, which are both pleasingly close to measured values, in particular in the case of $\dot{m}_{R,cut}^*$. These findings have several important implications. Firstly, the evolution of $v_{e,R}^*$ with $\dot{m}_R^*$ seems to support the linear correlation between $\dot{m}_R^*$ and $L_R^*$ proposed in Eq. \ref{eq:LR_mR}, including for what concerns the existence of a non null, minimum recirculation length $L_{R,0}^*$. Secondly, Eq. \ref{eq:ve_Uinf} and Eq. \ref{eq:ve_Uinf_2} might provide interesting insight into the physical significance of both parameters of Eq. \ref{eq:LR_mR}. In particular, the proportionality constant $k_{m,L}$ does not depend on actuation parameters, nor on properties determined by the incoming boundary layer, such as $\Rey_\theta$ or $\delta_e/h$, because $k_{m,L} \sim \left(v_{e,H}^*\right)^{-1}$ and $v_{e,H}^*\approx 0.02$ in both the present experiment and the unperturbed flow analysed in \citet{stella2017}. Since the two works share the same ramp profile and have similar expansion ratios, it seems more likely that $k_{m,L}$ mostly depends on geometry. As for what concerns $L_{R,0}^*$, it is tempting to relate its value to the frequency response of the actuators. Indeed, Eq. \ref{eq:LR_mR} yields: \begin{equation} L_{R,cut}^* = L_R^*\left(\dot{m}_{R,cut}^*\right)= 2 L_{R,0}^* \approx 1.6 \approx L_{R,min}^*, \end{equation} and hence $L_{R,0}^* \sim L_{R,min}^*$. Now, curves reported in Figure \ref{fig:Freq_resp} and Figure \ref{fig:LR_StA} suggest that $L_R^* \rightarrow L_{R,min}^*$ is broadly correlated to $U_{JP}^* \rightarrow U_{JP,max}^*$, i.e. that $L_{R,min}^*$ might be mainly determined by the saturation of the synthetic jets. Then, although at present no element allows to definitely discard an additional dependency on geometry, $L_{R,0}^*$ might be rather influenced by properties of the actuators. Finally, it seems important to remind that, at least to a certain extent, $U_{JP}^*$ can be considered anticorrelated to $\dot{m}_R^*$ (see Figure \ref{fig:Freq_resp}). This means that it should be possible to transpose Eq. \ref{eq:ve_Uinf_2} into a \textit{low-pass} filter that applies to the velocity of the synthetic jet. In other words, the control action becomes ineffective at reducing $L_R^*$ if $U_{JP}^*$ exceeds a value corresponding to $\dot{m}_{R,cut}^*$. Most importantly, for a given actuator this threshold seems to mainly depend on the geometry of the ramp. \begin{figure} \centering \includegraphics[width=\linewidth]{./fig14} \caption{Evolution of $v_{e,R}^*$ with $\dot{m}_R^*$. The grayscale indicates the value of $St_A$. The characteristic of the high-pass filter defined by Eq. \ref{eq:ve_Uinf_2} is shown in red online (\protect{\color{red}{\solidrule}}) and its linearised form in fine, black online (\protect\finesolidrule). Fine dashed lines (\protect\finedashedrule) highlight $v_{e,R}^*/v_{e,H}^*=0.5$ and $\dot{m}_R^* = L_R^*/k_{l,m}$. The decreasing trend of $v_{e,R}^*$ for $\dot{m}_R^* < 0.03$ is represented by a thick, dashed line (\protect\dashedrule), which was obtained by fitting a model in the form $v_{e,R}^*/v_{e,H}^*=k_{filt}\,\dot{m}_R^*$, where $k_{filt}$ is a constant, onto available data. The intersection of this linear fit with $v_{e,R}^*/v_{e,H}^*=0.5$ gives $\dot{m}_{R,cut}^* \approx 0.0123$.} \label{fig:ve_mR} \end{figure} \FloatBarrier \bibliography{2018_SJMK} \end{document} }}\end{equation}}
\caption{Evolution of $C_{p,b}$ with respect to $\dot{m}_R^*$. The grayscale indicates the value of $St_A$ and the dashed line (\protect\dashedrule) represents a linear fit.}
\caption{Evolution of $v_{e,R}^*$ with $\dot{m}_R^*$. The grayscale indicates the value of $St_A$. The characteristic of the high-pass filter defined by Eq. \ref{eq:ve_Uinf_2} is shown in red online (\protect{\color{red}{\solidrule}}) and its linearised form in fine, black online (\protect\finesolidrule). Fine dashed lines (\protect\finedashedrule) highlight $v_{e,R}^*/v_{e,H}^*=0.5$ and $\dot{m}_R^* = L_R^*/k_{l,m}$. The decreasing trend of $v_{e,R}^*$ for $\dot{m}_R^* < 0.03$ is represented by a thick, dashed line (\protect\dashedrule), which was obtained by fitting a model in the form $v_{e,R}^*/v_{e,H}^*=k_{filt}\,\dot{m}_R^*$, where $k_{filt}$ is a constant, onto available data. The intersection of this linear fit with $v_{e,R}^*/v_{e,H}^*=0.5$ gives $\dot{m}_{R,cut}^* \approx 0.0123$.}
\caption{ \textbf{Quantitative comparison with state-of-the-art methods.} % The methods with a $\star$ sign are trained on our LR-GOPRO training set. % {\color{red}Red texts} indicate the best performance. % The proposed GFN performs favorably against existing methods while maintaining a small model size and fast inference speed. }
\caption{\footnotesize Embodied agents that actively explore novel objects (left) or $360^{\circ}$ environments (right) intelligently select camera motions to gain as much information as possible with very few glimpses. While they naturally face \textcolor{Blue}{limited observability} of the environment, during \emph{learning} \textcolor{Red}{fuller observability} may be available. We propose sidekicks to guide policy learning for active visual exploration.}
\caption{Avg/Adv MSE errors $\times 1000$ ($\downarrow$ lower is better) and corresponding improvements (\%) over the \texttt{one-view} model ($\uparrow$ higher is better), for the two datasets. The best and second best performing models are highlighted in \textcolor{Green}{green} and \textcolor{Blue}{blue} respectively. Standard errors range from 0.2 to 0.3 on SUN360 and 0.1 to 0.2 on ModelNet Hard. (* - requires full observability at test time)}
\caption{Experimental measurements. (a) Representative stability diagram (differential conductance through the quantum point contact (QPC) as a function of gate voltages $V_\text{L}$ and $V_\text{R}$) for devices being considered, showing typical double dot behavior. (b) Representative stability diagram showing depletion of the double dot device to low electron occupations. \textcolor{red}{The (0,0)? indicates that the occupations are consistent with being zero, but the presence of a filled shell cannot be ruled out definitively.} (c) \textcolor{red}{Representative differential conductance through the QPC measured as a function of the baseline voltage applied to gate L, $V_\mathrm{L}$, when a square wave pulse is added to the dc baseline applied to gate L. This excited state spectroscopy is used to measure the singlet-triplet splitting in a dot, which is extracted from the difference in voltage of the lines indicated by the two arrows.} For these measurements, $V_{\mathrm{L}}$ is attenuated by a factor of 10, and the pulse height is attenuated by a factor of 33. (d) A representative magnetospectroscopy plot demonstrating Zeeman splitting of an effectively single-electron quantum dot via differential conductance through the quantum point contact (QPC) as a function of the applied magnetic field $B$ and the voltage on gate L, $V_\mathrm{L}$. The gate lever arms were determined using the relation that the Zeeman energy is equivalent to \SI{0.12}{\milli\electronvolt\per\tesla}. Adding a second electron to the quantum dot becomes less energetically favorable as the magnetic field increases, over the entire range measured, providing strong evidence that the singlet-triplet splitting in the dot is large. \label{fig:experimental_data}}
\caption{WEKA's hyperparemter tuning tool from \href{https://www.cs.ubc.ca/labs/beta/Projects/autoweka/}{UBC}. Note that central to the tool is a single hyperparameter optimizer called SMAC~\cite{hutter11smac}, shown in \textcolor{red}{{\bf RED}}. This paper asks the question ``is one hyperparameter optimizer enough?''. }
\caption{Flux and the diffusion equation. (a) Concentration profiles, $C(x,t)$, at times $t=3$ \tikzcircle[fill=green]{3pt} and $t=3.2$ s \tikzcircle[fill=black!60!green]{3pt}. (b) The flux of cells past a given point, $J$ (black; left axis), and the concentration gradient, $\partial C/\partial x$ (yellow; right axis), versus position, $x$. (c) Flux, $J$, versus concentration gradient, $\partial C/\partial x$, for all the values of $x$ and $t$ shown in \FIG{fig2}\refstyle{a}. The dashed magenta line has a slope $D=0.1$ mm$^2$/s.}
\caption{Examples of learned \textcolor{blue}{MTLUs} in FSRnet$_7$. MTLU$_{ij}$ denotes the activation function for the $j$-th channel of layer $i$. ReLU function (in \textcolor{red}{red}) is included for reference.}
\caption{%Weakly-supervised active learning via motion priors. Cross-domain human segmentation performance (IoU) comparison of the proposed weakly-supervised active learning method ``PAL'' with other strategies. \highlight{U- and Seg- denote the model architectures: U-Net and SegNet, respectively.} First row ``Source Only'' is direct application of pre-trained model on target domain data. To best of our knowledge, none of the existing active learning algorithm use only prior instead of true label for fine-tuning on target domain.}
\caption{\red{Energy level diagrams and experimental setup.} (a) Relevant molecular potential energy curves. Depicted at the right is a scattering state of a pair of atoms (the tri-colored spheres) whose spin part of their quantum state is a superposition of different bare $m_f$ spin states. Beneath is the decomposition of the scattering state as that of various pairs of atoms (the mono-colored spheres) with bare $m_f$ spin states. The superposition coefficients are denoted $C_{-1},C_{0},$ and $C_{+1}$; red, blue, and green represent $m_f=$ -1, 0, and +1, respectively. The non-zero CG coefficients for the $\ket{F=0, m_F=0}$ component of the various pairs of bare spins in the scattering state are shown near the thin black arrows. (b) Laser geometry. Two Raman lasers with angular frequencies $\omega_R+\Delta\omega_R$ and $\omega_R$ propagate along $\pm\hat{y}$ and have linear polarizations along $\hat{x}$ and $\hat{z}$. The frequency difference between the Raman beams $\Delta\omega_R/2\pi$ is $ 3.5$ MHz. A PA laser propagates in the x-z plane (due to spatial constraints) and has a linear polarization with components along all three axes. (c) Atomic energy diagram. A Zeeman bias magnetic field, $|\vec{B}_\mathrm{Bias}|\approx 5$ G is used to tune the Raman detuning $\delta.$ (d) Dressed band structures for Raman coupling $\Omega_R= 0,\,1.1,\,3.2, 8.0$ and $12\,E_r$, all with $\delta=0$. Dots represent BECs adiabatically loaded to the band minima at $q=0$. Panels (a) to (c) are not to scale.}
\caption{A Blockitecture\textregistered\set. The first five pictures show the process to build a garden tower, while the last picture shows the possibility of a different configuration.{\em Animation in supplemental video.}}
\caption{(a)--(b): Strongly convex $f$; dashed black curve is $O(1/k)$. (c)--(d): Nonconvex $f$; dashed black curve is $O(1/\sqrt{k})$. Color-coded curves denote: {\color{red} $\|w_k - w^*\|^2$}, {\color{cyan} $f(\bar{w}_k) - f(w^*)$}, {\color{blue} $f(w_k) - f(w^*)$}, and {\color{magenta} $\min_k\|\nabla f(w_k)\|^2$}.}
\caption{ 2AFC scores (higher is better) for different methods using disparity in deep feature representations \cite{zhang2018unreasonable} and texture representations (ours) on BAPPS validation dataset. Values in blue are \textcolor{blue}{highest performing} while the values in red are \textcolor{red}{the second best}. Our texture based scores from untrained supervised networks consistently perform better than the feature based scores and compare to *LPIPS metrics which are specifically trained on BAPPS training dataset, thus have an advantage over other untrained methods.}
\caption{\textcolor{red}{(a)} Our model contains three GANs representing three domains, i.e. homography view($X$), Bird view($Y$) and Frontal view($Z$). Each of them consists of three modules: Encoder($E$), Generator($G$) and Discriminator($D$). Since all three domains have a shared semantic meaning, we employ shared layers for learning unified intermediate representations. $G$s learn the cross-domain translations and reconstruction mappings based on the intermediate representation. $D$s are adversarial discriminators for the respective domains, distinguishing between real and fake samples. \textcolor{red}{(b)} Single cycle-consistency loss includes two components, i.e. forward cycle: $X\xrightarrow[]{\mathcal{G}} \widetilde{Y}^{x\to y} \xrightarrow[]{\hat{\mathcal{G}}} \widetilde{X}^{y\to x} \approx X$ and backward cycle: $Y\xrightarrow[]{\hat{\mathcal{G}}} \widetilde{X}^{y\to x} \xrightarrow[]{{\mathcal{G}}} \widetilde{Y}^{x\to y} \approx Y$. \textcolor{red}{(c)} To enforce the consistent cross-view features, we introduce the loss network (ResNet-50) for regularizing the inconsistent problem.}
\caption{Label Synchronous Viterbi Beam Search for CTC \textcolor[rgb]{0,0.5,0}{(Inputs: start and end nodes, token queue, time frames)}}
\caption{ Auto-encoding capabilities of the model with errors marked in \textcolor{red}{\textbf{bold red}}. The model was trained only on English Wikipedia paragraphs. On short sequences, our model performs close to an identity function. On longer ones, it seems to correctly auto-encode only English paragraphs. Note that the model tries to map French words into English ones (\texttt{avec} $\rightarrow$ \texttt{open}, \texttt{une} $\rightarrow$ \texttt{one}). We observed a similar behavior on other languages as well. }
\caption{Color hierarchy (best to worst) is \colorbox{RoyalBlue!90}{blue}, \colorbox{RoyalBlue!35}{light blue}, \colorbox{red!45}{light red}, \colorbox{BrickRed}{red}, for each network \& metric. Across all networks, BTER is usually first or second for degree\& CC distributions, yet usually poor in modularity; EGBTER is usually best or second in all three metrics. All metrics are macro-averaged across 100 random realizations for each model.}
\caption{Comparison of quantitative results in terms of PSNR and SSIM on four synthesized benchmark datasets. The three best performing algorithms are marked in \color[HTML]{FE0000}\textbf{red}\color{black}, \color[HTML]{3166FF}\textbf{blue}\color{black}, and \color[HTML]{32CB00}\textbf{green}\color{black}, respectively. Our proposed NLEDN consistently achieves the best performance. $`*'$ indicates that the method uses additional data~(e.g. rain density level, rain mask annotation) provided by the dataset.}
\caption{MAE(lower is better) and F-measure(higher is better) comparisons with 7 methods on four open datasets. The top three results are highlighted in {\color{red}red}, {\color{green}green}, and {\color{blue}blue} fonts respectively.}
\caption{Comparison of the dispersion of the energetically lowest energy-loss peak based on EELS measurements ($\times$) and BSE simulations (\textcolor{red}{\large$\circ$}) for \qq{} parallel to the (a) $\Gamma$K and (b) $\Gamma$M directions. The gray dotted line in (a) represents the EMA fit of the EELS data.}
\caption{(Color online) The first and second columns show the IP transition maps for the valence and conduction states, respectively, for the energetically lowest exciton ($\lambda=1$) and the indicated momentum transfer values parallel to the $\Gamma$M direction in the $\Gamma$KM plane. The black dots outlining the Brillouin zone represent the \textit{k} points used in the calculations where log$(\abs{\sum_{v,c}M^{\mathbf{k}vc}_\lambda(\bm{q})}/\abs{I_\lambda(\bm{q})})$\,=\,0. The gray lines trace the high symmetry lines. The orientation and length of the black arrows beginning at the center of the plots symbolize the direction and magnitude of the momentum transfers employed in the computations. The plots in the third column depict the DFT band structure after applying the scissor operator and the IP transitions derived from the IP transition maps for the indicated momentum transfer values parallel to$\Gamma$M. The insets show the same transitions in the $\Gamma$KM plane of the Brillouin zone. }
\caption{Energy-momentum dispersions of the IP transitions ({\textcolor{red}{\large$\circ$}}) and the energetically lowest exciton ({\large$\times$}) for \qq{} parallel to the (a) $\Gamma$K and (b) $\Gamma$M directions. All points have been blue-shifted by 1.4~eV.}
\caption{\label{fig:cluster}The fraction of particles in the largest cluster, $n_{\mathrm{c}}$, for $\Delta U = 10 k_{\mathrm{B}}T$ at three $\phi$ (labelled), as a function of time $t$ (in Brownian time units $t_{\mathrm{B}}$): \textcolor{red}{$\boldsymbol{\boxdot}$} = NH, \textcolor{blue}{$\boldsymbol{\odot}$} = WH; $\boldsymbol{\times}$ (WH) and $\boldsymbol{+}$ (NH) indicate percolation points; and error bars give standard errors. Black dotted lines and numbers indicate the exponent of asymptotic power laws, with standard error $\pm 0.1$.}
\caption{\label{fig:voids}Structural properties for gelling systems with $\Delta U = 10 k_{\mathrm{B}}T$, (\textcolor{red}{$\boldsymbol{\boxdot}$} = NH, \textcolor{blue}{$\boldsymbol{\odot}$} = WH), and for three $\phi$ as indicated by the labels and use of line type; the result for $\Delta U = 5 k_\mathrm{B}T$ is analogous. (a) Void volume (VV) probability density functions (PDFs) at the aging time $t_{\mathrm{a}}$. Error bars indicate the standard error, which is typically smaller than the symbol size. One in five data points is shown to improve the presentation, but the lines that serve as guides to the eye go through every data point. (b) The time dependence of the void growth length, $\lambda_{\mathrm{VV}}$, see Eq.~\eqref{eq:lambda}, again only one in five data points is shown. ($\boldsymbol{+}$,$\boldsymbol{\times}$) and ($\blacksquare$,$\square$) = percolation ($t_{\mathrm{p}}$) and aging onset times, respectively. The number in parentheses indicates the factor by which $\lambda_{\mathrm{VV}}$ is multiplied to separate the data. The thin dashed black lines and numbers indicate power-law fits and associated exponents (with error $\pm 0.005$), respectively. Only for $\phi = 0.075$ is there an appreciable difference in the NH/WH initial exponent. (c) Structure factors $S(q)$ at $t_{\mathrm{a}}$, corresponding to the data in (a). We only show error bars up to $q \sigma = 1$ for every third data point to improve the presentation, the curves connect all data points; the region of the central peak is not shown $q \sigma < 0.2$. (d) The peak position of the structure factor $q_{\mathrm{m}}$ as a function of time. The three WH/NH data sets have been shifted up by the amount indicated in parentheses to prevent overlaps. The dashed line indicates a power law with exponent $-0.04 \pm 0.005$.}
\caption{Structure's response to step-sine excitation around its first natural frequency. (a, b) Markers (\textcolor{orange}{$\boldsymbol{\diamondsuit}$}, \textcolor{dgreen}{$\boldsymbol{+}$}, \textcolor{purple}{$\boldsymbol{*}$}, \textcolor{dred}{$\boldsymbol{\Box}$}) correspond to tests performed using identical parameters. The test performed between 10-15 Hz (\textcolor{dblue}{$\boldsymbol{\bullet}$}) serves to illustrate the overall shape of the resonance region. Branch of isolated periodic responses is captured only for the two tests (\textcolor{orange}{$\boldsymbol{\diamondsuit}$}, \textcolor{purple}{$\boldsymbol{*}$}) presented in Subfigure (b). (c) Quasi-periodic oscillations observed at 12.6 Hz, i.e. between the resonance peak and the isolated solution. (d) Poincar\'e map of the time series presented in Subfigure (c).}
\caption{Response curves measured at 13.5 Hz exhibit a S-like shape. Markers (\textcolor{dyellow}{$\boldsymbol{+}$}, \textcolor{dorange}{$\boldsymbol{+}$}, \textcolor{dblue}{$\boldsymbol{*}$}) correspond to different runs of the same test and serve to demonstrate the repeatability of the experimental results.}
\caption{Response of the beam structure as a function of the harmonic excitation amplitude and frequency. (\textcolor{black}{$\boldsymbol{\bullet}$}) Experimentally-measured data point. (\textcolor{dblue}{$\boldsymbol{-}$}, \textcolor{dorange}{$\boldsymbol{-}$}, \textcolor{dgreen}{$\boldsymbol{-}$}, \textcolor{dred}{$\boldsymbol{-}$}) Curve of constant force amplitude obtained using regression and numerical continuation.}
\caption{Verification of the invasiveness of the control signal for one of the S-curves presented in Figure~\ref{fig:Scurve_repeat}. (a) RMS value of the higher-harmonic component (\textcolor{orange}{$\boldsymbol{-\bullet-}$}) and the non-harmonic component (\textcolor{blue}{$\boldsymbol{-\diamond-}$}) of the control signal relative to its fundamental component. (b) Absolute RMS value of the higher-harmonic component (\textcolor{orange}{$\boldsymbol{-\bullet-}$}) and the non-harmonic component (\textcolor{blue}{$\boldsymbol{-\diamond-}$}) of the control signal.}
\caption{Verification of the invasiveness of the control signal. Histogram of the RMS value of the higher-harmonic component (\textcolor{orange}{$\blacksquare$}) and the non-harmonic component (\textcolor{dblue}{$\blacksquare$}) of the control signal for all the data points collected.}
\caption{Effect of the higher harmonics in the applied force on the dynamic behaviour of the system for the lower (a, c) and the upper (b, d) periodic solutions highlighted in Figure~\ref{fig:stab_1p5N}(a). (a, b) Relative importance of the higher harmonics compared to the fundamental component of the force applied to the system. (c, d) Phase portraits of the steady-state periodic orbits. Closed-loop experimental conditions before (\textcolor{dblue}{$\blacksquare$}) and after (\textcolor{orange}{$\blacksquare$}) applying the harmonic compensation procedure. Open-loop experimental conditions (\textcolor{dyellow}{$\blacksquare$}).}
\caption{In a series of experiments, we compare two sets of binary codes constructed with two different affinity matrices: class {(a)} and instance based {(b)}. {(c)-(e)} contrasts the binary codes with respect to residual norm, $\mathsf{mAP}$ and inference time. Results for binary codes inferred from the class and instance affinity matrices are denoted with (\blueline) and (\protect \redlined), respectively. We also learn hash functions with varying complexities to fit the inferred binary codes and plot the fraction of non-matched bits to the total number bits {(f)}.}
\caption{Norm of the residual matrix \vs iteration, when step sizes in Alg. Binary code inference are \texttt{constant} (\blueline) and \texttt{regress} (\redline). Notice the residual norm of the \texttt{regress} inference scheme converges both faster and to zero.}
\caption{Example of an RRC. The task was to predict if $A$ is \pred{AdvisedBy} $B$, given the relations of people at a university.}
\caption{ \textbf{Qualitatively different steady-state degree distributions for different $q$.} a), b) For small $q$ ($q=0.1$ in the examples) the stationary degree distributions are peaked at $k=\left<k\right>-1 $ whereas c), d), for larger $q$ ($q=0.5$ in the example), they become monotonically decreasing in $k$, thus peaking for $k=0$. The transition is discontinuous (jumping from peak position $\left<k\right>$ to zero) and robust against changes of initial conditions (Erdős-Rényi initial networks, panels a and c; Barabási-Albert initial networks, panels b and d). Example ensembles shown for initial networks of $N=2^{12}$ nodes and $\left<k\right>=20$. Darker colors indicate longer times and gray wide lines indicate the analytical prediction as time $t\rightarrow \infty$ (stationary distribution). Data are averaged across 100 realizations of the stochastic temporal network evolution; error bars show standard deviation of resulting distribution. Note the different vertical scales in panel d) compared to c).} \label{fig.sim} \end{figure} Here, we study a class of network restructuring processes with explicit \emph{adhesion} preference of a link to stay connected to its end nodes. The resulting networks exhibit a genuine discontinuous phase transition in their macroscopic structure as an adhesion parameter changes. A degree-dependent asymmetry of the node-link adhesion induces a transition between classes of networks with qualitatively different degree distributions, one monotonically decaying with degree, the other peaked. Intriguingly, the two classes consistently distinguish abstract (e.g. online) from face-to-face (e.g. offline) social networks as we confirm by comparing analytical and simulational results of our theory with structural network data for empirical systems. \textit{Degree-dependent network restructuring.} Consider a basic restructuring process of a non-growing temporal network of $N$ nodes and $L$ links, starting from an arbitrary interaction topology. In each time step $t\in\mathbb{N}_0$ of restructuring (Fig.\,\ref{fig.process}), a link in the network is chosen at random from the uniform distribution among all links. One of the two end nodes of that link is cut (`given up' by the unit it connected to), with probability $q$ choosing the lower-degree node and with probability $1-q$ the higher-degree node. The cut end of the link reconnects to a different unit of some degree $k$, randomly chosen with a probability distribution proportional to $(k+1)$. The resulting time evolution defines a stochastic ensemble of temporal networks. After sufficiently long times, we consistently observe convergence to a network ensemble dependent only on $q$ (unique attractor). Indeed, the degree distributions become stationary and independent of the initial network structure (Fig.\,\ref{fig.sim}). For low $q$ the stationary distribution is peaked, for higher $q$ it is decaying. Below, we evaluate the order parameter \begin{equation} m = \evdel{k}^{-1} \mathrm{argmax}[\tilde{P}(k)] \label{eq.mk} \end{equation} given by the normalized mode (position of the maximum) of the stationary degree distribution $\tilde{P}(k)$ and show that the system exhibits a well-defined phase transition in the thermodynamic limit where $\evdel{k} \rightarrow \infty$ and thus $N \rightarrow \infty$. To qualitatively understand how the transition emerges and how it depends on adhesion and network topology, we derive and analyze a master equation characterizing the degree distribution $P_t(k)$ of the evolving network ensemble. Consider the possible degree changes at a given time step $t\geq 0$. The one end node that sticks at the randomly chosen link does not change its degree, whereas the other end node reduces its degree by one. Subsequently, the node that the link rewires to increases its degree by one. As a consequence, the degree distribution of the temporal network process satisfies the discrete-time master equation \begin{align} P_{t+1}(k)&=P_t(k)-u_k P_t(k)+u_{k+1}P_t(k+1)\nonumber \\ &-l_k P_t(k)+l_{k-1}P_t(k-1). \label{eq.rec2} \end{align} The unwiring probability $u_k$ represents the probability of decreasing the degree of one of the nodes with degree $k$ by rewiring away from it and the linking probability $l_k$ the probability of increasing the degree of a node with degree $k$ by rewiring to it. As the temporal network evolution is a Markov chain that is irreducible, i.e. every network in the ensemble can be reached with positive probability, the stationary degree distribution ${\tilde{P}}(k)$ is unique and given as solution of Eq.\,\ref{eq.rec2} with $P_{t+1}(k)=P_t(k)=\tilde{P}(k)$. We rewrite Eq.\,\ref{eq.rec2} at the fixed point as a matrix equation \begin{equation} \begin{aligned} \underbrace{ \begin{pmatrix} \text{-}(l_0) & u_1 & 0 &\cdots & 0\\ l_0 & \text{-}(l_1+u_1) & u_2 &\cdots & 0 \\ \vdots &\vdots & \ddots &\vdots& \vdots \\ 0 & \cdots&0 & l_{N\text{-}2} & \text{-}u_{N\text{-}1} \end{pmatrix}}_{\mathbf{M}} \begin{pmatrix} {\tilde{P}}(0)\\ {\tilde{P}}(1) \\ \vdots \\ {\tilde{P}}(N\text{-}1) \end{pmatrix}=0. \end{aligned} \label{eq.fix2} \end{equation} where ${\tilde{P}}$ is a vector with entries ${\tilde{P}}(k)$ and $\mathbf{M}$ a matrix with entries based on $u_k$ and $l_k$. Adding to each row of the matrix $\mathbf{M}$ the previous row and dividing by $u_i$ simplifies Eq.\,\ref{eq.fix2} to \begin{align} \underbrace{ \begin{pmatrix} \frac{l_{0}}{u_1} & -1 & 0 &\cdots & 0\\ 0 & \frac{l_{1}}{u_2} & -1 &\cdots & 0 \\ \vdots &\vdots & \ddots &\vdots& \vdots \\ 0 & \cdots&0 & 0 & 0 \end{pmatrix}}_{\mathbf{M}'} \begin{pmatrix} {\tilde{P}}(0)\\ {\tilde{P}}(1) \\ \vdots \\ {\tilde{P}}(N-1) \end{pmatrix}=0. \label{eq.fix3} \end{align} Since the last row of $\mathbf{M}'$ is identically zero, $\mathbf{M}'$ does not have full rank and there exists a stationary solution for $k \in \{1,\ldots,N-1\}$, given by the eigenvector to the eigenvalue zero as \begin{align} p(k)&= \left(\prod_{m=1}^{k}\frac{l_{m-1}}{u_m}\right). \label{eq.Pgen} \end{align} We fix $p(0)$ arbitrarily and then normalize $\tilde{P}(k)=Cp(k)$ by $C=1/\sum_{k=0}^{N-1} {p}(k)$ for all k such that $\sum_{k=0}^{N-1} {\tilde{P}}(k)=1$ to obtain the exact stationary degree distribution. \begin{figure}[!th] \centering \includegraphics[width=\columnwidth]{transition} \caption{\textbf{Phase transition in temporal network rewiring}. a) The relative position of the maximum $m =\evdel{k}^{-1} \mathrm{argmax}[\tilde{P}(k)]$ of the stationary degree distribution $\tilde{P}(k)$ as a function of adhesion asymmetry parameter $q$ for different network sizes with $N > \evdel{k} \gg 1$, where $\evdel{k}/N = \mathrm{const.}$ %b) The position of the transition point $q_c( \evdel{k}/2 )$ where $m = 1/2$ is algebraically approaching $1/2$ in the thermodynamic limit. b) The distance $\epsilon(k)$ from the transition point (\ref{eq.epsilon}) algebraically decays to zero in the thermodynamic limit for all $k\leq \evdel{k}$, see the examples for $k/\evdel{k}=\{0,1/2,1\}$, thus indicating a discontinuous transition at $q_c = 1/2$. The scaling $\epsilon(k) \sim \evdel{k}^{-1}$ agrees with the analytical approximation (\ref{eq.eps}) and the exact prediction (\ref{eq.eps0}). c) Phase diagram in parameter space $q$ vs. $\evdel{k}$ indicates separation between decaying ($m = 0$) and peaked ($m > 0$) degree distributions (white/shaded regions, resp.). The analytical prediction (solid line) shows $q_c(0)$ from Eqns.~(\ref{eq.eps0}) and (\ref{eq.epsilon}), disks show the result from direct numerical simulations for $N = 1024$.} \label{fig.pt} \end{figure} \textit{Asymmetric adhesion induces discontinuous transitions.} To calculate the peak position $k_\mathrm{max}=\mathrm{argmax}[\tilde{P}(k)]$ of the degree distribution, we first consider the unlinking and linking probabilities $u_k$ and $l_k$. Since links are selected uniformly at random, the probability of a link with an end node of degree $k$ being selected is $k/L$, where $L = N\evdel{k}/2$ is the number of links in the network. That node is chosen to be cut from the link with probability $q$ if $k$ is smaller than the degree of the other end node of the link and with probability $1-q$ if it is larger than that degree. For uncorrelated degrees of the two end nodes we obtain the approximation \begin{align} u_k&=\frac{k}{L}\left(q P^{\scriptscriptstyle>}(k)+ \frac{1}{2} P^{\scriptscriptstyle=}(k) + (1-q) P^{\scriptscriptstyle<}(k)\right), \label{eq.pu} \end{align} where $P^{\scriptscriptstyle>}(k) = \sum_{k'=k+1}^{N-1} \frac{k'}{\evdel{k}}\tilde{P}(k')$, $P^{\scriptscriptstyle=}(k) = \frac{k}{\evdel{k}}\tilde{P}(k)$ and $P^{\scriptscriptstyle<}(k)=\sum_{k'=1}^{k-1}\frac{k'}{\evdel{k}}\tilde{P}(k')$ are the probabilities that a node neighbouring a degree $k$ node has itself a degree larger than, equal to, or smaller than $k$. Numerical results suggest that for networks where $\evdel{k}$ is substantially smaller than $N$ the node degrees are indeed sufficiently weakly correlated in the stationary state for $q \le 1/2$. The cut link rewires following preferential attachment \cite{Barabasi1999} and rewires to a new node with probability proportional to $(k+1)$. The offset of $1$ prevents a node from being removed if its degree falls to zero. So the (linking) probability of reconnecting to a node of degree $k$ is \begin{equation} l_k = \frac{k+1}{\sum_{i=1}^{N}(k_i+1)}=\frac{k+1}{N\evdel{k+1}}. \label{eq.pl} \end{equation} \begin{figure*}[t] \centering \includegraphics[width=16 cm]{dataletters} \caption{\textbf{Social networks in the phase diagram.} a) The phase transition (dark blue line) theoretically derived from the simple network restucturing model separates all personal (dark blue) from all abstract (light green) social networks studied. The topologies of 33 different networks were extracted from data collected in \cite{Kunegis2013}. For each network $i$, the optimal parameter $q_i$ given the average degree $\evdel{k}_i$ (and network size $N_i$) as extracted from the network data, are placed in the phase diagram. The inset magnifies a densely covered area of high school social networks. b-e) Examples of social networks illustrating decaying and peaked degree distributions, the solid curves indicate single-parameter fit of the entire distributions. b) Online community ``Hamsterster'' \cite{konect:2016:petster-friendships-hamster}, c) University Rovina email exchange \cite{konect:guimera03}, d) interactions in an exhibition \cite{konect:sociopatterns}, e) and friendships among school children \cite{konect:moody}.} \label{fig.data} \end{figure*} To understand how the peak position changes, we calculate the values of $q=:q_c(k)$ where the maximum of the degree distribution changes from $k$ to $k+1$ in the steady state ensemble. We then demonstrate that $q_c(k) \rightarrow 1/2$ for all $k\leq \evdel{k}$ in the limit as $\evdel{k}\rightarrow \infty$ and thus $N\rightarrow \infty$ such that the order parameter discontinuously jumps at that value of $q$. Fig.~\ref{fig.pt} displays finite system results and illustrates the scaling as $\evdel{k}$ and $N$ grow. For finite systems, the point $q_c(k)$ is defined by the condition that the two consecutive probabilities become equal, $\tilde{P}(k) = \tilde{P}(k+1)$. Using $P^{\scriptscriptstyle>}(k) + P^{\scriptscriptstyle=}(k) + P^{\scriptscriptstyle<}(k) = 1$ and writing the probabilities $P^{\scriptscriptstyle>}(k) = (1 - P^{\scriptscriptstyle=}(k))/2 + \delta^{\scriptscriptstyle>}(k)$ and $P^{\scriptscriptstyle<}(k) = (1 - P^{\scriptscriptstyle=}(k))/2 + \delta^{\scriptscriptstyle<}(k)$, where $\delta^{\scriptscriptstyle>}(k) = - \delta^{\scriptscriptstyle<}(k)$, Eqs.~(\ref{eq.Pgen} - \ref{eq.pl}) yield the equation \begin{align} 1 &= \frac{\tilde{P}(k)}{\tilde{P}(k+1)} = \frac{u_{k+1}}{l_k} \nonumber \\ &=\frac{2\evdel{k+1}}{\evdel{k}} \left[ \frac{1}{2} + \left(2q_c(k)-1\right) \delta^{\scriptscriptstyle>}(k) \right] \label{eq.delta} \end{align} for $q_c(k)$. For $q_c = 1/2$ the right-hand-side is larger than one for all finite $\evdel{k}$. We thus consider the deviation \begin{align} \epsilon(k) = 1/2 - q_c(k) \geq 0 \,, \label{eq.epsilon} \end{align} such that (\ref{eq.delta}) results in \begin{align} \epsilon(k) &= \frac{1}{4 \evdel{k+1} \delta^{\scriptscriptstyle>}(k)} \,. \label{eq.eps} \end{align} This provides strong evidence for a discontinuous phase transition, because in the limit $\evdel{k} \rightarrow \infty$ we have $\epsilon(k)\rightarrow 0$ for all $k$ as long as $\delta^{\scriptscriptstyle>}(k) > 0$. We make additional progress by noting that $\delta^{\scriptscriptstyle>}(k)$ changes sign at the median $\bar{k}$ of the distribution. For the typical (though not universal) ordering of the mode $k_\mathrm{max}$, median $\bar{k}$ and mean $\evdel{k}$ of a unimodal distribution with positive skewness \cite{groeneveld1977mode,abadir2005mean}, $k_\mathrm{max} \le \bar{k} \le \evdel{k}$, we thus find that indeed $\delta^{\scriptscriptstyle>}(k_\mathrm{max}) > 0$ and $q_c(k) < 1/2$, consistent with the transitions observed in the simulations of finite systems (Fig.~\ref{fig.pt}a). As a result, the peak position changes discontinuously from $m=\frac{k_\mathrm{max}}{\evdel{k}} = 0$ to $m=1$ at $q_c = 1/2$ in the limit of $\evdel{k} \rightarrow \infty$ and thus $N\rightarrow \infty$. For $k=0$ we specifically have $P^{\scriptscriptstyle>}(0) = 1$ and $P^{\scriptscriptstyle=}(0) = 0$ and consequently $\delta^{\scriptscriptstyle>}(0) = 1/2$ such that we obtain the exact expression \begin{align} \epsilon(0) &= \frac{1}{2 \evdel{k+1}} \,. \label{eq.eps0} \end{align} Direct numerical simulations show that the same scaling holds for $\epsilon(k)$ across values of $k \in \{0,...,\evdel{k}\}$, compare Fig.~\ref{fig.pt}b The approximate scaling form (\ref{eq.eps}), the exact result (\ref{eq.eps0}) and the numerical analysis summarized in Fig.~\ref{fig.pt} jointly indicate a discontinuous phase transition at $q_c=1/2$ in the limit $\evdel{k} \rightarrow \infty$, and thus $N\rightarrow \infty$. We emphasize that the order parameter changes abruptly from $m = 1$ for all $q<1/2$ to $m = 0$ for $q>1/2$ as the adhesion parameter $q$ continuously varies across it's critical value from below. \textit{Transition line separates abstract from personal social networks.} Interestingly, a wide range of social networks exhibit one of these two specific types of degree distributions -- decaying or peaked -- theoretically identified above. Moreover, the theoretical transition line exactly separates the collection of networks into those with %face-to-face or otherwise close personal contacts and those created through more abstract, indirect or online-only interactions, compare Fig.\,\ref{fig.data}: We have compared the degree distributions resulting from the simple model % analyzed in this article to those obtained from 33 social networks, as reported in references \cite{Kunegis2013} and \cite{starbuck,konect:boguna,konect:dolphins,konect:coleman1957,konect:ucidata-zachary,konect:arenas-jazz,konect:freeman1998,konect:knuth1993,konect:boguna,DBLP:Massa}. The systems range from networks of direct personal interactions with face-to-face contacts in various private and educational contexts or reported friendships of humans and of dolphins, to more abstract social interactions, including online social networks as well as offline, but hierarchically %, not individually determined social relations (see supplemental material for more details). The average degree $\left<k\right>$ was computed from each data set thus leaving $q$ as the only free parameter. A least-squares fit to each degree distribution yields the best-suitable $q$ for each network. Intriguingly this even led to good quantitative agreement of the degree distributions (see Fig.\,\ref{fig.data}\,b-e,\blue{see also supplemental material}). %, especially considering the often small number of data points. %Beyond this quantitative finding, we note that strict qualitative finding of decaying vs. peaked degree distributions found, to our knowledge a fact that has not been noted before. We were amazed to observe that all networks in which personal, typically face-to-face relationships define links exhibit a peaked degree distribution, whereas all networks in which abstract, i.e. online relationships define links, exhibit decaying degree distributions, see Fig.\,\ref{fig.data}. Indeed, a permutation test assigning the labels 'abstract' and 'personal' to the degree distributions randomly yields the same classification only in one out of $\binom{33}{8}\approx 1.4\times 10^7$ cases. %We find even quantitative agreement in properties of the distributions: all observed peaks are consistent with the predicted positions $\evdel{k}$ and $0$ for the peaked and decaying distributions, respectively and the fitted curves are in reasonable agreement, see Fig.\,\ref{fig.data}\,b-e.\\ \textit{Discussion and Conclusions.} Studies of network structure forming processes have previously uncovered explosive transitions in a range of network growth processes \cite{DSouza2015} whereas most restructuring processes for fixed size networks %that do not grow, including the Watts-Strogatz model \cite{Watts1998}, exhibit gradual cross-overs between random and regular graphs with no distinct transition. Notably, the rewiring mechanisms in the latter types of processes are independent of any properties of the nodes or links \cite{Watts1998, grosskinsky2002universal,grabow2012small,watts1999small,newman2000mean,Molkenthin2016a}. Here, we revealed a discontinuous phase transition in network structural features in a simple model class of temporal networks that do not grow but exhibit degree-dependent link adhesion. As the underlying microscopic mechanisms rely on a simple local cutting and rewiring process, they imply self-organization of the networks' large-scale structures. Driven by the tendency of a link to stick to a node (adhesion) depending on that node's degree, a discontinuous phase transition emerges between two types of degree distributions -- peaked and decaying -- when smoothly varying this tendency via a control parameter. Intriguingly, the same two types of degree distributions are observed for a wide range of social networks. Moreover, among the 33 topologies of social networks analyzed, all those networks established through personal contacts exhibit peaked and all those with more abstract, indirect or impersonal contacts exhibit decaying degree distributions. The results thus suggest that the simple, abstract model, that is \emph{a priori} unrelated to specific and widely heterogeneous social dynamics, surprisingly serves as a good indicator for the separation between networks with more personal and more abstract social relations. Several general constraints as well as social mechanisms might be supporting this binary classification. The primary mechanisms underlying the model restructuring process are based on the model ingredient that sustaining a link is node-dependent. From the perspective of a node in a socio-economic setting, such as trade or friendships, this may be interpreted as the %level of requiring bilateral effort (e.g., time, money or motivation) to keep a connection. Such constraints on the number of sustainable links may lead a node to cut ties with less beneficial connections (i.e. a less involved friend or a less influential business partner), similar to our model settings that otherwise are far from fully capturing the intricacies across social networks. %For instance, the more effort an interaction requires, the more a social or economic agent might tend to keep the number of its interaction partners low and potentially cut links. %Such a constraint about overall effort, likely higher for personal than for abstract networks, may serve as one option potentially supporting the emergence of distinct classes of social networks. Taken together, the results presented above not only highlight severe theoretical consequences of link adhesion -- inducing a discontinuous transition for restructuring networks in the first place -- and yield novel insights into structural phase transitions in temporally evolving networks \cite{HOLME201297,PhysRevLett.110.198701,scholtes2014causality}. They may also serve as a starting point for future investigations about the mechanisms and the influence of constraints in evolving socio-economic systems. \textit{Acknowledgements.} We thank Stefan Grosskinsky, Yael Fender, Frank Schweitzer, Alex Arenas and Johanna Wolter for valuable comments on presentations about this work. This work is supported through the German Science Foundation (DFG) by a grant towards the Cluster of Excellence `Center for Advancing Electronics Dresden' (cfaed). \bibliographystyle{unsrt} \bibliography{socialnets.bib} \end{document}}
\caption{ The architecture of Joint Sequence Fusion (JSFusion) model. %The left part (a) of architecture is Joint Semantic Tensor (JST) and the right part (b) is Convolutional Hierarchical Decoder (CHD). {\bf\color{MidnightBlue}{Blue paths}} indicate the information flows for multimodal similarity matching tasks, while {\bf\color{OliveGreen}{green paths}} for the fill-in-the-blank task. (a) JST composes pairwise joint representation of language and video sequences into a 3D tensor, using a soft-attention mechanism. (b) CHD learns hierarchical relation patterns between the sequences, using a series of convolutional decoding module which shares parameters for each stage. $\odot$ is Hadamard product, $\oplus$ is addition, and $\otimes$ is multiplication between representation and attentions described in Eq.(\ref{eq:jst})--(\ref{eq:convatt}). We omit some fully-connected layers for visualization purpose. }
\caption{Comparison of top \textbf{Laptop} responses generated for different scenarios by adaptation training VRALSTM (denoted by $\flat$) and VDANLG (denoted by ${\sharp}$) models from Source domains, and by training VRALSTM from \textit{\textbf{Scratch}}. Errors are marked in colors (\textcolor{red}{[missing]}, \textcolor{violet}{misplaced}, \textcolor{orange}{redundant}, \textcolor{teal}{wrong}, \textcolor{olive}{spelling mistake} information). \textcolor{blue}{[OK]} denotes successful generation. VDANLG$^\sharp$ = VRALSTM$^\flat$+SC+DC.}
\caption{Comparison of top responses generated for different scenarios by adaptation training VRALSTM (denoted by $\flat$) and VDANLG (denoted by ${\sharp}$) models from Source domains, and by training VRALSTM from \textit{\textbf{Scratch}}. Errors are marked in colors (\textcolor{red}{[missing]}, \textcolor{violet}{misplaced}, \textcolor{orange}{redundant}, \textcolor{teal}{wrong}, \textcolor{olive}{spelling mistake} information). \textcolor{blue}{[OK]} denotes successful generation. VDANLG$^\sharp$ = VRALSTM$^\flat$+SC+DC.}
\caption{\label{fig:model} The outcome of the modification model. The first three networks that originally demonstrate presence of RCC, i.e., AS, Bible and Software get transformed to networks with no RCC. The last three networks that originally do not demonstrate the presence of RCC, i.e., Power, Protein and Facebook get converted to networks with RCC. These plots are similar to the eigen gap chart of Figure~\ref{fig:fig_shell}, except we show the eigengap over all the shells rather than in groups. The \textcolor{blue}{blue} (\textcolor{green}{green}) plot shows the eigengap for the original (modified) network.}
\caption{\label{fig:model} The outcome of the modification model. The first three networks (top panel) that originally demonstrate presence of RCC, i.e., AS, Bible and Software get transformed to networks with no RCC. The last three networks (bottom panel) that originally do not demonstrate the presence of RCC, i.e., Power, Protein and Facebook get converted to networks with RCC. These plots are similar to the eigen gap chart of Figure~\ref{fig:shell_fig}(c), except we show the eigengap over all the shells rather than in groups. The \textcolor{blue}{blue} (\textcolor{green}{green}) plot shows the eigengap for the original (modified) network.}
\caption{\label{tab:jacs} Jaccard index between nodes with highest coreness and equal number of high centrality nodes. Results clearly separate the two categories of networks into ones that have an RCC (\textcolor{blue}{blue}) and ones that do not have an RCC (\textcolor{brown}{brown}).}
\caption{Histograms of \textcolor{mycolor7}{\textbf{minimum}}, \textcolor{mycolor8}{\textbf{maximum}}, and \textcolor{mycolor9}{\textbf{mean}} instantaneous speed (m/s) for each window in the labeled modes of locomotion.}
\caption{Speeds from an independent test sequence featuring movement on three floor levels. The background color indicates the motion mode (\textcolor{mycolor11}{\textbf{static}}, \textcolor{mycolor12}{\textbf{walking}}, and \textcolor{mycolor10}{\textbf{stairs}}).}
\caption{We avoid predicting depth at the target pose, i.e., $P\to$ {\color{red}{depth prediction network}} $\to O$, by the following depth completion strategy: $P\to$ warp to target pose $\to \tilde{O}\to$ {\color{green}{depth completion network}} $\to O$.}
\caption{Qualitative results of depth completion methods at different viewing location and orientation on NYU Depth v2~\cite{Silberman:ECCV12} dataset. Images are first warped to the target pose and then use depth completion methods to predict depth at the occluded regions. {\color{green}{Depth-Flow-Net}} produces less artifacts and can even hallucinate handles of the chairs. A complete video is shown in the supplementary material.}
\caption{Correlation analysis of two adjacent pixels for, (a)-(c) Airplane (d)-(f) Serrano. \textcolor{blue}{$\circ$} and \textcolor{red}{$\circ$} relate to the original and encrypted images, respectively.}
\caption{\label{fig:8}(Color online) Theoretical temperature rise of water (red solid line), and a solution of gold nanomatryoshkas ({\color{red}dark} green solid line) with geometry [$r_1,r_2,r_3$]=[25.5,38.4,65] nm {\color{red}and a solution of 65-nm TiN nanoparticles (orange solid line)} as a function of time. Dotted, dashed and dashed-dotted lines correspond to experimental data \cite{22} of the temperature rise of water, TiN and carbon solution having a concentration of 0.0001 vol $\%$, respectively. {\color{red}Inset: Temperature increase as a function of time due to a contribution of TiN nanoparticles calculated using theory (dark green) and experiment (orange).}}
\caption{CDF of ventilation \\ duration by race (\textcolor{red}{$p=.005$}). }
\caption{\textbf{Mechanical Ventilation}\\ \textbf{High Trust}: 4810 patients\\ \textbf{Low Trust}: \;\;\;510 patients\\ \textcolor{red}{ $p<0.001$ }}
\caption{\textbf{Vasopressors}\\ \textbf{High Trust}: 4456 patients\\ \textbf{Low Trust}: \;\;\;453 patients\\ \textcolor{red}{p=0.001}}
\caption{\textbf{Mechanical Ventilation}\\ \textbf{High Trust}: 4810 patients\\ \textbf{Low Trust}: \;\;\;510 patients\\ \textcolor{red}{p<0.001}}
\caption{\textbf{Mechanical Ventilation}\\ \textbf{High Trust}: 4646 patients\\ \textbf{Low Trust}: \;\;\;492 patients\\ \textcolor{red}{$p<0.001$}}
\caption{ Left: { \it ROSAT} image of the Vela SNR (G263.9$-$3.3). Soft emission is shown in red and hard emission in blue/green. The contour is the outer radio boundary of Vela X. The individual circles identify \xmm\pointings discussed in the text. The dashed circular region was taken in small-window mode and was used for imaging, but not for spectral analysis. The numbered circles in the lower right illustrate regions used for spectral fitting of each pointing (see Figure 6). Upper right: Zoomed-in region of{\it ROSAT} image showing pulsar and cocoon. The outermost radio contour is shown to illustrate the extent of Vela~X. Lower right: \chandra\image of the Vela pulsar and its surrounding compact nebula. The jet axis is in the SE-NW direction, roughly aligned with the pulsar proper motion. The direction to the Vela~X cocoon is indicated.}
\caption{ Left: MOST radio image of Vela X showing filamentary structure within the extended nebula. The single contour is the outermost contour from the diffuse radio emission in Vela X. White regions in the image correspond to negative fluxes associated with missing coverage due to a lack of small baselines. Right: Exposure-corrected mosaic of \xmm\pointing in Vela~X, shown in blue, along with the MOST radio image, shown in red. The bright, elongated radio structure in the central region lies adjacent to the X-ray cocoon.%Radio image provided by A. Hales. }
\caption{ Maps of the power law index (left), temperature in keV (middle), and Ne abundance relative to the ISM value (right) from spectral fits of subregions in \xmm\pointings in Vela~X. The temperature and abundance correspond to the ejecta component in the spectral model. The outermost radio contours provide the rough outline of the PWN and the cross indicates the position of the pulsar. The color bars indicate the values for the photon index, the temperature in keV, and the abundance relative to solar values, respectively. The parameter uncertainties vary from region to region, but are typically$\lsim 0.05 - 0.1$ for the spectral index, $\sim 0.01$~keV for the temperature, and $\sim 0.1 - 0.3$ for the abundance. Distinct spectral steepening is observed along the cocoon, which also shows higher temperatures and abundances than other regions within the PWN. For reference, region 7 for each of the pointings (see inset to Figure 1, left) is identified with a red dot in the center panel. Dashed circles correspond to regions for which spectra are shown in Figure 7. }
\caption{ SNR radius evolution for expansion in a cloudy ISM (solid), and the Sedov solution (dashed) for evolution into a uniform medium, both for $E_{51} = 1$. The cloudy ISM curves (from Slavin et al. 2017) assume density values of $0.25 (25) {\rm\ cm}^{-3}$ for the intercloud (cloud) regions. The value of $C$ corresponds to the ratio of the cloud mass to that of the intercloud medium. The Sedov solution curves provide reasonable approximations assuming densities roughly connected to the evaporated cloud density in the {\red cloudy ISM} model. }
\caption{ Forward (red) and reverse (blue) shock evolution with time using solutions of Truelove \& McKee (1999). The solid (dashed) curves correspond to ambient densities of$n_0 = 0.15 (0.5) {\rm\ cm}^{-3}$ with $E_{51} = 1$, $M_{ej} = 8 M_\odot$, and a power law index $n = 12$ for the outer ejecta. }
\caption{Quantitative evaluation of state-of-the-art SR algorithms. \textbf{{\color{red} Red}} and {\color{blue} \underline{blue}} indicate the best and the second best performance, respectively.}
\caption{Diagnostic quality assessment in terms of subjective quality scores for different algorithms (mean$\pm$stds).~\textbf{{\color{red} Red}} and {\color{blue}{blue}} indicate the best and the second best performance, respectively.}
\caption{Simulation of 10000 cells at various time steps with a single cell infected initially, which is located at (0.5,0.5). The states are denoted as \textcolor{black}{$\blacksquare$} Susceptible, \textcolor{red}{$\blacksquare$} Infected, \textcolor{green}{$\blacksquare$} Recovered. The simulation parameters are defined as: contact tolerance - $\kappa=0.95$, infectivity radius - $\rho_0=0.04$, infection time - $T_I = 30$, recovered time - $T_R = 30$, and spatial step - $\Delta r = 0.001$. As time increases, the epidemic spreads as a wave throughout the domain. The inset on (a) shows the radially symmetric neighborhoods of a few infected cells.}
\caption{\textbf{Evaluating $\bm{\hat{V}_d(\pi_e)}$} for each reward component $d$, across policies for four labs. For randomized policies, the error bars show the standard deviation across ten trials. The ({\color{Maroon}$\star$}) indicates the best performing policy for each reward component.}
\caption{ Hysteresis on directed random graphs of connectivity $z=2,3$ for a Gaussian distribution N(0,$\sigma^2$) of the random field. We have set $J=1$. Theoretical results are shown by continuous curves. Symbols depict results from numerical simulations of the model {\color{blue}{for a single configuration of the random-field distribution on a system of size $N=10^6$. Numerical results are indistinguishable from the theoretical results on the scale of the figure}}. Hysteresis is absent on a $z=2$ graph for any $\sigma$ and also on a $z=3$ graph if $\sigma > \sigma_c \approx 1.781$. For $\sigma < \sigma_c$ the loop has discontinuities at $h=\pm h_c$. The discontinuities reduce in size and move towards $h=0$ with increasing $\sigma$. As $\sigma \to \sigma_c$, the discontinuities vanish continuously at $h_c=0$.}
\caption{ Hysteresis on a partially directed random graph with $z=3$ and $\sigma=1.5$ in units of $J$. A fraction $c$ of the nodes have two directed edges and one undirected. All three edges of the remaining fraction $1-c$ are undirected. Theoretical results are shown by continuous lines and simulations {\color{blue}{on a $N=10^6$ graph}} by symbols. Vertical portions of curves denote discontinuities. Results shown in Fig.1 are recovered for $c=0$. As $c \to 1$, the first-order jump in $m(h)$ reduces and moves away from the origin. A case where all three edges are undirected is also shown for comparison.}
\caption{Comparison of $\delta g_b(E)$ %\eq{dgb} obtained with the Bessel averaged $P(n)$ (red) with the coarse-grained trace formula $\delta g_\gamma(E)$ %\eq{dgscga} (blue). The amplitudes were adjusted by a constant factor. Note that the two curves are out of phase by $\delta E \sim 0.5$ (see text). {\color{red} (Should we show this?)} }
\caption{\red{Fractional rms amplitude of the type-C QPOs as a function of photon energy. The black, blue and red points represent LE, ME and HE data respectively.}\label{fig:f8}}
\caption{\red{Centroid frequency of the type-C QPOs as a function of energy. The black, blue and red points represent LE, ME and HE data respectively.}\label{fig:f6}}
\caption{ALMA Band 10 images of the integrated intensity of the redshifted high velocity gas ($-3$ to +2.5 \kms\/; the systemic velocity is$-7$~\kms\/) for the (a) ground state of HDO and the (b) CS (18-17) transition showing a highly collimated north-south outflow emanating from MM1. Blueshifted high velocity emission (not shown) is co-spatial with the redshifted emission, but is much weaker. The 0.35~mm continuum is the same as Fig.~\ref{map}. White circles show the locations of H$_2$O masers detected with the VLA in epoch 2017.8 \citep{Brogan18}. Primary beam correction has been applied with a cutoff at 0.25 the FWHM. The synthesized beam of $0\farcs23\times 0\farcs16$ (PA$=39\arcdeg$) is shown in the lower right of each panel.}
\caption{Simulated spectra of \ce{HC(O)CH2OH} plotted in {\color{red}\textbf{red}} over ALMA observations of NGC\,6334I MM1 plotted in\textbf{black}. The simulated spectra assume $T_{\rm{ex}}$~=~135~K, $\Delta V$~=~3.2~\kms\/, and$N_T$~=~$1.3\times10^{17}$~cm$^{-2}$, with a $v_{lsr}$~=~-7~\kms\/. The ALMA observations have been convolved to a uniform synthesized beam of 0.26\arcsec$\times$0.26\arcsec. In the lower six panels, smaller regions of the frequency coverage in Band 10 have been selected to show detail. }
\caption{Processing flow and intermediate data tensors of CBinfer. Color code: \textcolor{babyblueeyes}{custom processing kernels}, \textcolor{green!60!black!60!}{cuBLAS kernel}, \textcolor{yellow!20!orange!80!}{variables sharable among layers}, and \textcolor{red!40!orange!60!}{variables to be stored per-layer}. Size and data type of intermediate results are indicated below the variable name. Coarse-grained feed-forward CBinfer (a) is introduced in \secref{sec:procSteps}, the closed-loop formulation (b) is described in \secref{sec:closedLoop}, and the fine-grained extensions to the algorithm (c,d) are formulated in \secref{sec:fineGrained}.}
\caption{\change{Evaluation of accuracy effect, runtime, number of compute operations, power and energy when performing inference for the pose detection network on a 200-frame video sequence.} \textbf{Legend:} {\color{red}CBinfer}, {\color{blue}cuDNN}; continuous: max-N (max. performance) power mode, dashed: max-Q (max. efficiency) power mode.}
\caption[font=small]{The illustration of StepGAN. The generator's encoder (\textcolor{red}{red}) reads in sentence $x$, and then the generator's decoder (\textcolor{BurntOrange}{orange}) produces output sentence $\hat{y}$. The discriminator (\textcolor{blue}{blue} and \textcolor{LimeGreen}{green}) either reads in the generated sentence pair $(x,\hat{y})$ or the target sentence pair $(x, y^*)$, and then outputs estimated Q-values $\hat{Q}(s_t,\hat{y}_{t})$ (or $\hat{Q}(s_t,{y}^*_{t})$) at each time step $t$. The average of estimated Q-values at every time step $t$ is taken as the final discriminator score $D(x,\hat{y})$ (or $D(x,y^*)$).}
\caption{The broadband SED of \bl{} for the active (left) and low VHE \gray{} emitting states. The boosted synchrotron/SSC emission from the jet are shown in gray whereas the \gray{} spectra from {\it pp} interactions and synchrotron emission from secondary $e^{-}e^{+}$ pairs are shown with blue color. The data are corrected for EBL absorption adopting the model from \citet{dominguez}.}
\caption{The tracking results of video \emph{person7} in out-of-view challenge. First row: tracking snapshots of SiamRPN and DaSiamRPN. Second row: detection scores and according overlaps of the two methods. The overlaps are defined as intersection-over-union (IOU) between tracking results and ground truth. \color{red} Red: ground truth. \color{green} Green: tracking box. \color{blue} Blue: Search region box.}
\caption{Performance comparisons on public short-term benchmarks. OP: mean overlap precision at the threshold of 0.5; DP: mean distance precision of 20 pixels; EAO: expected average overlap, and mean speed (FPS). The {\color{red}\textbf{red bold}} fonts and {\color{blue}\textit{blue italic}} fonts indicate the best and the second best performance.}
\caption{Illustration of how three DYNK functions are stored in memory. The colours are used to separate the three functions used in the example ({\color{red} f1}, {\color{blue} f2}, {\color{brown} f3}). The function {\color{blue} f1} is of type FILE, where the column Data1 points to where in \texttt{fexpr\_dynk} the data begins, and Data2 indicates how much data there is. The function {\color{blue}f2} is of type LIN, where Data1 points to where in \texttt{fexpr\_dynk} the two parameters are stored. The function {\color{brown}f3} is of type ADD, where Data1 and Data2 indicates which functions should be added by listing their row numbers in \texttt{funcs\_dynk}.}
\caption{Example Markov blanket $MB_T$ of a target $T$, consisting of three parents (\textcolor{mambacolor4}{blue} nodes), two children (\textcolor{mambacolor3}{green}) and one spouse (\textcolor{mambacolor5}{orange}). All other nodes are $\independent T$ given $MB_T$. Our goal is to discover the MB, \emph{as well as} the edge directions.}
\caption{Substitution results w.r.t. human expert selection \emph{distribution} and ERSATZ \emph{similarity} responses. Note that, gray cells correspond to object categories which are not available in the respective query, cells marked with \tikzdrawcircle[black, fill=black]{2.5pt} represents substitutes selected by experts and ERSATZ.}
\caption{Solution of the Helmholtz equation for the most critical wavenumber. ----- Analytic solution. \fullcirc, standard mesh; \fullsquare, optimized mesh.\label{fig:mesh}}
\caption{Stability for each combination of algorithm variant and corpus. Measured with j@10 metric (higher is better). Same data as in Table \ref{tab:results-short}, standard deviations too small to \textcolor{red}{display}.\label{fig:rel}}
\caption{Illustration of latency components for DL (\textcolor[RGB]{10,89,178}{blue}) and UL (\textcolor[RGB]{227,130,65}{orange}) transmissions. The latency components are defined in Table~\ref{assumptions}.}
\caption{Evaluation on three benchmarks in comparison to the state-of-the-art unsupervised video re-id methods. {\color{red} Red}: the best performance. {\color{blue} Blue}: the second best performance. `-': no reported results.}
\caption{Effectiveness of two association losses. {\color{red} Red}: the best performance. CNN: MobileNet.}
\caption{Comparison with supervised counterparts. {\color{red} Red}: the best performance. CNN: MobileNet.}
\caption{SQL query generation in SQL-B. Syntax errors are highlighted in {\color{red}\textbf{red}} color.}
\caption{Python code generation in Django. Syntax errors are highlighted in {\color{red} \textbf{red}} color.}
\caption{Spectrum of the ``diffuse'' emissions toward the GC measured by the HESS instrument \citep{2016Natur.531..476H} together with absorption-corrected data (note only absorption-corrected points $>10$~TeV are shown and these are offset compared to the original data by -10\% (R12) and +10\% (F98) in energy, respectively, for clarity). Point styles and colours: black open, uncorrected; red solid triangle, R12; cyan solid square, F98. Lines show the \gray{} spectral fit models used to estimate the 95\% lower confidence level to the proton cutoff energies (see text). Line types: solid, no ISRF correction; short-dashed, R12; long-dashed, F98. \label{fig:hessgc}}
\caption{Top: Ratio of the PIYs of multiply charged ions M$^{q+}$ and single charged ions M$^{+}$ for naphthalene as a function of the photon energy. Bottom: Comparison of our result for the anthracene molecule at photon energy of 2500 eV with the \textcolor{blue}{Postma et al. (2010)} results using proton and $\alpha$-particle impacts.}
\caption{ Two examples of lattices. (A) A one-dimensional lattice with properties $d = 1$, $\mathcal{N} = \{0,-1,1\}$ and edge weights $\psi(0) = p$ and $\psi(-1) = \psi(1) = s$. This is the generic form of the one-dimensional nearest neighbor lattice we cover in more detail in section 3. (B) A two-dimensional lattice with properties $d = 2$, $\mathcal{N} = \left\{ (0,0),(0,1),(1,0),(-1,-1) \right\}$, and edge weights $\psi(0,0) = p$, $\psi(0,1) = s_1$, $\psi(1,0) = s_2$, $\psi(-1,-1) = s_3$. } \label{fig:lattice} \end{figure} % Two examples of lattices are shown in Fig. \ref{fig:lattice}. The first example in Fig. \ref{fig:lattice}(A) is a one-dimensional nearest neighbor lattice. Here, each node $v_{\textbf{i}}$ has $|\mathcal{N}|=3$ neighbors; $v_{\textbf{i}-1}$, $v_{\textbf{i}+1}$, and the node itself, $v_{\textbf{i}}$ for every $\textbf{i} \in \mathcal{Z}$. The second example in Fig. \ref{fig:lattice}(B) is a two-dimensional lattice with directed edges listed in the caption. %Here, each node $v_{\textbf{i}}$ is connected $|\mathcal{N}| = 4$ neighbors, namely, $\mathcal{N} = \{(0,0),(0,1),(1,0),(-1,-1)\}$. We emphasize that the remaining results in this section are applicable to any lattice of arbitrary dimension and connectivity pattern that can be described by Definition \ref{def:lattice}.\\ % \indent We also find it useful to define some additional operations on integer vectors in order to maintain clarity in some of the following equations. % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % OPERATIONS ON INTEGER TUPLES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \begin{defn}{(Operations on Integer Tuples)} % \begin{itemize} % \item Let the product of two indices be defined elementwise, that is, the $k$th element of the product $(\textbf{i} \textbf{j})_k = i_k j_k$. % \item Let the Dirac delta function between two indices be defined as $\delta_{\textbf{i},\textbf{j}} = \prod_{k=1}^d \delta_{i_k,j_k}$ where $\delta_{i_k,j_k} = 1$ if $i_k = j_k$ and $\delta_{i_k,j_k} = 0$ otherwise. % \item Let the exponential of an index be $e^{a \textbf{i}} = \prod_{k=1}^d e^{a i_k}$ where $a$ is a complex coefficient. % \item Let the integral with respect to an integer vector be defined as, % \begin{equation} \int_a^b f(\textbf{i}) d\textbf{i} = \int_a^b \cdots \int_a^b f(\textbf{i}) d i_1 \ldots d i_d \end{equation} \end{itemize} \end{defn} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % LINEAR DYNAMICS % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \subsection{Linear Dynamics} % Each node $v_{\textbf{i}}$ in our lattice $\mathcal{L}$ is endowed with a time-varying state $x_{\textbf{i}}(t) \in \mathcal{R}$. These time-varying states may represent position or velocity for formation type problems \cite{yu2010some}, or transfer rates in routing problems \cite{zhao2005onset}, or average excitement in neuronal networks \cite{wang2015passivity}, and many other network applications. Before presenting the dynamical system that describes the time evolution of the states of each node, we will introduce two finite subsets of the nodes in the lattice. % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % DEFINITION 2: DRIVER NODES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \begin{defn}[Driver Nodes] Let $\mathcal{D} \subset \mathcal{V}$ be the set of driver nodes of cardinality $|\mathcal{D}| = n_d$. If node $v_{\textbf{i}} \in \mathcal{D}$, then node $v_{\textbf{i}}$ receives an external control input $u_{\textbf{i}}(t)$ which we are free to define. We assume that no control input can be connected to more than a single node. \end{defn} % %\begin{rem} %A control input $u_{\textbf{i}}(t)$ can be attached to only a single driver node $v_{\textbf{i}}$, so that no two driver nodes can share the same control input. %Often in networked systems there is physical distance between the nodes and so allowing multiple nodes access to the same control input may not be physically realistic. %\end{rem} % \begin{defn}{(Target Nodes)} Let $\mathcal{T} \subset \mathcal{V}$ be the set of target nodes of cardinality $|\mathcal{T}| = n_t$. If $v_{\textbf{i}} \in \mathcal{T}$, then there is a desired value for the state $x_{\textbf{i}}(t_f) = x_{\textbf{i},f}$ which is chosen before applying the control action. \end{defn} % \begin{rem} % Note that we will assume both $\mathcal{D}$ and $\mathcal{T}$ are finite subsets of $\mathcal{V}$ and they may overlap, that is, it is possible for a node $v_{\textbf{i}} \in \mathcal{D} \cup \mathcal{T}$. \end{rem} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % The linear differential equation that governs the time evolution of the each state is, % \begin{equation}\label{eq:sys} \dot{x}_{\textbf{i}}(t) = \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) x_{\textbf{i}+\textbf{n}}(t) + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} u_{\textbf{a}}(t). \end{equation} % While here we consider scalar dynamics ($x_\textbf{i}(t) \in \mathcal{R}$ and $\psi(\textbf{n}) \in \mathcal{R}$), the extension to multi-dimensional states, $\textbf{x}_{\textbf{i}}(t) \in \mathcal{R}^{n_x}$ and $\boldsymbol{\psi}(\textbf{n}) \in \mathcal{R}^{n_x \times n_x}$ is straightforward. % Note that the time evolution of the states of each node $v_{\textbf{i}}$ are a function of its immediate neighbors $v_{\textbf{i}+\textbf{n}}$ for all $\textbf{n} \in \mathcal{N}$ and possibly a driver node if $v_{\textbf{i}} \in \mathcal{D}$. Each node has an initial condition $x_{\textbf{i}}(0) = x_{\textbf{i},0}$ for all $v_{\textbf{i}} \in \mathcal{V}$, and the target nodes have a final condition $x_{\textbf{i}}(t_f) = x_{\textbf{i},f}$ for all $v_{\textbf{i}} \in \mathcal{T}$. We would like to achieve this final condition with minimal control energy. % \begin{equation}\label{eq:opt} \begin{aligned} \min && &J = \frac{1}{2} \int_0^{t_f} \sum_{v_{\textbf{a}} \in \mathcal{D}} u_{\textbf{a}}^2(\tau) d\tau\\ \text{s.t.} && &\dot{x}_{\textbf{i}}(t) = \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) x_{\textbf{i}+\textbf{n}}(t) + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} u_{\textbf{a}}(t), \quad v_{\textbf{i}} \in \mathcal{V}\\ && &x_{\textbf{i}}(0) = x_{\textbf{i},0}, \quad v_{\textbf{i}} \in \mathcal{V}, \quad x_{\textbf{i}}(t_f) = x_{\textbf{i},f}, \quad v_{\textbf{i}} \in \mathcal{T} \end{aligned} \end{equation} % The optimal cost of this optimal control problem can be written \cite{klickstein2017energy}, % \begin{equation}\label{eq:optcost} J^* = \frac{1}{2} \textbf{b}^T \bar{W}^{-1} \textbf{b} \end{equation} % The vector $\textbf{b}$ has entries representing the difference between the prescribed final condition $x_{\textbf{i},f}$ and what the state would be without any external control input. The matrix $\bar{W}$ is the output controllability Gramian \cite{klickstein2017energy} which is a principal submatrix of the controllability Gramian corresponding to the indices of the target nodes. The elements of the controllability Gramian obey the Lyapunov equation, % \begin{equation}\label{eq:Wdot} \begin{aligned} \dot{W}_{\textbf{i},\textbf{j}}(t) &= \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) W_{\textbf{i}+\textbf{n},\textbf{j}}(t) + \sum_{\textbf{n}\in\mathcal{N}} \psi(\textbf{n}) W_{\textbf{i},\textbf{j}+\textbf{n}}(t)\\ &+ \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}}. \end{aligned} \end{equation} % The solution of Eq. \eqref{eq:Wdot} for systems described by lattice graphs is derived by applying the $2d$-dimension discrete time Fourier transform. % \begin{thm} Let $\mathcal{I} = \sqrt{-1}$, the unit complex value, and let, % \begin{equation} \phi(\textbf{k}) = \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) e^{-\mathcal{I} \textbf{n} \textbf{k}} \end{equation} % where we call $\phi(\cdot) : \mathcal{Z}^d \mapsto \mathcal{C}$ the lattice function for lattice $\mathcal{L}$. The time evolution of the controllability Gramian entries for the infinite lattice is, % \begin{equation}\label{eq:Wt} \begin{aligned} W_{\textbf{i},\textbf{j}}(t) &= \frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} e^{-\mathcal{I}(\textbf{i}-\textbf{a}) \hat{\textbf{i}}} e^{-\mathcal{I} (\textbf{j}-\textbf{a}) \hat{\textbf{j}}}\\ &\times \theta(t,\hat{\textbf{i}},\hat{\textbf{j}}) d\hat{\textbf{i}} d\hat{\textbf{j}} \end{aligned} \end{equation} % where the time varying portion, % \begin{equation} \begin{aligned} \theta(t,\hat{\textbf{i}},\hat{\textbf{j}}) %\int_0^t e^{\left(\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}})\right)t} dt\\ = \frac{\exp{\left[\left(\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}})\right)t\right]} - 1}{\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}})} \end{aligned} \end{equation} % \end{thm} % \begin{pf*}{Proof.} First, we state two facts: \begin{itemize} \item Let $a$ be a complex number. Then $a \int_0^t e^{a\tau} d\tau = e^{at} - 1$. \item Let $m$ be an integer. An important integral that appears is $\int_{-\pi}^{\pi} e^{-\mathcal{I} m x} dx = 2 \pi \delta_{m,0}$ % \end{itemize} % From Eq. \eqref{eq:Wt}, it is simple to show that, % \begin{equation}\label{eq:sumW} \begin{aligned} \sum_{\textbf{n}\in\mathcal{N}} \psi(\textbf{n}) W_{\textbf{i}+\textbf{n},\textbf{j}} &= \frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} \phi(\hat{\textbf{i}})\\ &\times e^{-\mathcal{I} (\textbf{i}-\textbf{a})\hat{\textbf{i}}} e^{-\mathcal{I}(\textbf{j}-\textbf{a}) \hat{\textbf{j}}} \theta(t,\hat{\textbf{i}},\hat{\textbf{j}}) d\hat{\textbf{i}}d\hat{\textbf{j}} \end{aligned} \end{equation} % Applying Eq. \eqref{eq:sumW} and the two facts stated above to Eq. \eqref{eq:Wdot}, we can complete the proof, % \begin{equation} \begin{aligned} &\dot{W}_{\textbf{i},\textbf{j}}(t) = %\frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} \left(\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}}) \right)\\ % %&\times e^{-\mathcal{I} (\textbf{i}-\textbf{a}) \hat{\textbf{i}}} e^{-\mathcal{I}(\textbf{j}-\textbf{a}) \hat{\textbf{j}}} \theta(t,\hat{\textbf{i}},\hat{\textbf{j}}) d\hat{\textbf{i}} d\hat{\textbf{j}} + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}}\\ % \frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} e^{-\mathcal{I}(\textbf{i}-\textbf{a})\hat{\textbf{i}}} e^{-\mathcal{I} (\textbf{j}-\textbf{a})\hat{\textbf{j}}}\\ % &\times\left( e^{(\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}}))t} - 1 \right) d\hat{\textbf{i}} d\hat{\textbf{j}} + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}}\\ % &= \dot{W}_{\textbf{i},\textbf{j}}(t) - \frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi}e^{-\mathcal{I}(\textbf{i}-\textbf{a}) \hat{\textbf{i}}}\\ % &\times e^{-\mathcal{I}(\textbf{j}-\textbf{a})\hat{\textbf{j}}} d\hat{\textbf{i}} d\hat{\textbf{j}} + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}}\\ % &= \dot{W}_{\textbf{i},\textbf{j}}(t) - \sum_{v_{\textbf{a}} \in \mathcal{D}} \prod_{k=1}^d \left[ \frac{1}{2\pi} \int_{-\pi}^{\pi} e^{-\mathcal{I}(i_k-a_k)\hat{i}_k} d\hat{i}_k \right]\\ % &\times\left[ \frac{1}{2\pi} \int_{-\pi}^{\pi} e^{-\mathcal{I}(j_k-a_k)\hat{j}_k} d\hat{j}_k \right] + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}}\\ &= \dot{W}_{\textbf{i},\textbf{j}}(t) - \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}} + \sum_{v_{\textbf{a}} \in \mathcal{D}} \delta_{\textbf{i},\textbf{a}} \delta_{\textbf{j},\textbf{a}} = \dot{W}_{\textbf{i},\textbf{j}}(t) \end{aligned} \end{equation} % \end{pf*} % \begin{rem} Of particular interest to us is the case when $\phi(\textbf{k}) < 0$ because then there exists a steady state solution to Eq. \eqref{eq:Wdot}, % \begin{equation}\label{eq:W} \begin{aligned} &W_{\textbf{i},\textbf{j}} = \lim_{t \rightarrow \infty} W_{\textbf{i},\textbf{j}}(t)\\ &= -\frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} \frac{e^{-\mathcal{I}(\textbf{i}-\textbf{a})\hat{\textbf{i}}} e^{-\mathcal{I}(\textbf{j}-\textbf{a})\hat{\textbf{j}}}}{\phi(\hat{\textbf{i}}) + \phi(\hat{\textbf{j}})} d\hat{\textbf{i}} d\hat{\textbf{j}}. \end{aligned} \end{equation} % When the dynamical system is stable, the minimum energy expression in Eq. \eqref{eq:optcost} approaches a constant value. Note that $\phi(\textbf{k}) < 0$ implies $\boldsymbol{0} \in \mathcal{N}$ and $\psi(\boldsymbol{0}) < \sum_{\textbf{n} \in \mathcal{N},\textbf{n} \neq \boldsymbol{0}} \psi(\textbf{n})$, i.e., there exists a large enough, and negative, self-loop at each node. \end{rem} % From Eq. \eqref{eq:Wdot}, we know $W_{\textbf{i},\textbf{j}}(t)$ is a real number and so the complex portion of both Eq. \eqref{eq:Wt} and \eqref{eq:W} must be zero. With this in mind, Eq. \eqref{eq:Wt} can be rewritten as, % \begin{equation}\label{eq:Wt_real} \begin{aligned} W_{\textbf{i},\textbf{j}}(t) &= \frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} \left[ \frac{(\alpha_{\textbf{a}} \sigma + \beta_{\textbf{a}} \omega)(r(t)-1)}{\sigma^2 + \omega^2} \right.\\ &\left.- \frac{(\beta_{\textbf{a}}\sigma - \alpha_{\textbf{a}}\omega) s(t)}{\sigma^2 + \omega^2} \right] d\hat{\textbf{i}} d\hat{\textbf{j}} \end{aligned} \end{equation} % where the functions are, \begin{equation} \begin{aligned} r(t) &= e^{\sigma t} \cos \omega t, \quad s(t) = e^{\sigma t} \sin \omega t\\ \sigma &= \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) \left[ \cos \left( \sum_{k=1}^d n_k \hat{i}_k \right) + \cos \left( \sum_{k=1}^d n_k \hat{j}_k \right) \right]\\ \omega &= \sum_{\textbf{n} \in \mathcal{N}} \psi(\textbf{n}) \left[ \sin \left( \sum_{k=1}^d n_k \hat{i}_k \right) + \sin \left( \sum_{k=1}^d n_k \hat{j}_k \right) \right]\\ \alpha_{\textbf{a}} &= \cos \left(\sum_{k=1}^d \left[ (i_k - a_k) \hat{i}_k + (j_k-a_k) \hat{j}_k \right]\right)\\ \beta_{\textbf{a}} &= \sin \left( \sum_{k=1}^d \left[ (i_k-a_k)\hat{i}_k + (j_k-a_k) \hat{j}_k \right] \right) \end{aligned} \end{equation} % Similarly, Eq. \eqref{eq:W} can be rewritten as, % \begin{equation}\label{eq:W_real} W_{\textbf{i},\textbf{j}} = -\frac{1}{(2\pi)^{2d}} \sum_{v_{\textbf{a}} \in \mathcal{D}} \int_{-\pi}^{\pi} \int_{-\pi}^{\pi} \frac{\alpha_{\textbf{a}} \sigma + \beta_{\textbf{a}} \omega}{\sigma^2 + \omega^2} d\hat{\textbf{i}} d\hat{\textbf{j}} \end{equation} % The expressions in Eqs. \eqref{eq:Wt_real} and \eqref{eq:W_real} lend themselves to multi-dimensional numerical integration as the imaginary portion has been removed. % \section{Nearest Neighbor Lattice} % %With Eqs. \eqref{eq:Wt} or \eqref{eq:W} (or equivalently Eqs. \eqref{eq:Wt_real} or \eqref{eq:W_real}) we can compute any element of the controllability Gramian. % % The one-dimensional ($d=1$), nearest neighbor ($\mathcal{N} = \{0,1,-1\}$) lattice (NNOD lattice) has the property that the shortest path between nodes $v_i$ and $v_j$ is $|i-j|$. An example of this lattice is shown in Fig. \ref{fig:lattice}(A). Let the edge weights be $\psi(1) = \psi(-1) = s > 0$ and $\psi(0) = p < -2s$. The lattice function of the NNOD lattice is $\phi(\hat{i}) = p + 2s \cos \hat{i} < 0$. As the Gramian entries are linear with respect to the contribution of each of the driver nodes, we examine each driver node separately so we assume $n_d = 1$. We shift the indices $i' = i-a$ and $j'=j-a$, to place the driver node at $a' = 0$. With this offset in mind, let $G_{i',j'} = W_{i,j}$ represent the shifted Gramian entries. % Equation \eqref{eq:W_real} for the NNOD lattice can be written as, % \begin{equation}\label{eq:Wnn} G_{i',j'} = \frac{-1}{\pi^2} \int_{0}^{\pi} \int_{0}^{\pi} \frac{\cos(i'\hat{i}) \cos(j'\hat{j})}{2p + 2s \cos \hat{i} + 2s \cos \hat{j}} d\hat{i} d\hat{j}. \end{equation} % An analogous version of the double integral in Eq. \eqref{eq:Wnn} has been derived separately in the context of solving the discretized Helmholtz equation \cite{katsura1971lattice,morita1971useful}.\\ % \indent \begin{lem}[Recursion for Diagonal Values \cite{morita1971useful}] Let $\alpha = p^2/(2s^2)-1$, which, if the system is stable corresponds to a value of $\alpha > 1$. The diagonal values, $G_{i',i'}$, can be found from the recursion relation, % \begin{equation}\label{eq:rr1} G_{i'+1,i'+1} = \frac{4i'}{2i'+1} \alpha G_{i',i'} - \frac{2i'-1}{2i'+1} G_{i'-1,i'-1} \end{equation} % with initial values, % \begin{equation}\label{eq:init} \begin{aligned} G_{0,0} &= \frac{1}{\pi |p|} K \left( \frac{4s^2}{p^2} \right)\\ G_{1,1} &= \left( \frac{|p|}{2 \pi s^2} - \frac{1}{\pi |p|} \right) K \left( \frac{4s^2}{p^2} \right) - \frac{|p|}{2\pi s^2} E \left( \frac{4s^2}{p^2} \right) \end{aligned} \end{equation} % where $K(\cdot)$ and $E(\cdot)$ are the first and second complete elliptic integrals, respectively. \end{lem} % \begin{cor} Let $z_{i'} = G_{i',i'} / G_{i'-1,i'-1}$ for $i' > 0$ represent the instantaneous rate of decay along the diagonal. The recursion in Eq. \eqref{eq:rr1} can be rewritten in terms of $z_{i'}$. % \begin{equation} z_{i'+1} z_{i'} = \frac{4i'}{2i'+1} \alpha z_{i'} - \frac{2i'-1}{2i'+1} \end{equation} % For large $i'$, this yields the approximate solution, % \begin{equation} \tilde{z} = \alpha - \sqrt{\alpha^2-1}, \end{equation} % so that the asymptotic rate of decay of the diagonal elements is exponential, % \begin{equation}\label{eq:scaling} G_{i',i'} \sim (\tilde{z})^{i'}. \end{equation} \end{cor} % \begin{figure} \centering \includegraphics[width=\columnwidth]{diag.pdf} \caption{Asymptotic behavior of the diagonal values of the controllability Gramian of the NNOD lattice. The marks represent the values computed by numerically integrating Eq. \eqref{eq:Wnn} using two dimensional Gauss quadrature and the dashed lines represent the scaling in Eq. \eqref{eq:scaling}. We see that even for small values of $\alpha$ the asymptotic behavior provides a good approximation for small values of the diagonal index $m$.} \label{fig:diag} \end{figure} % % In Fig. \ref{fig:diag} we show that for both large and small values of $\alpha$, Eq. \eqref{eq:scaling} does a satisfactory job approximating $G_{i',i'}$ even for small values of $i'$. % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % SUBSECTION: CONTROL METRICS % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Control Metrics} % %Note that the minimum control energy expression in Eq. \eqref{eq:optcost} is a function of the inverse of the output controllability Gramian. Two particularly useful control metrics are the trace of the inverse of the output Gramian and the logarithm of the determinant of the output Gramian \cite{summers2016submodularity}. We assume the system is output controllable, so that $\bar{W}$ is positive definite and order the eigenvalues $0 < \lambda_1 \leq \lambda_2 \leq \ldots \leq \lambda_{n_t}$. % As we will see next, with knowledge of the diagonal elements, we can lower bound the control metrics using the Cauchy interlacing theorem \cite{golub2012matrix}. % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % THEOREM: CAUCHY INTERLACING % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{thm}[Cauchy Interlacing Theorem \cite{golub2012matrix}] Let $X$ and $Y$ be two symmetric matrices of size $n$ and $m$ respectively, $n > m$, such that $Y$ is a principal submatrix of $X$. Order their eigenvalues such that $\lambda_i(X) \leq \lambda_{i+1}(X)$ and $\lambda_i(Y) \leq \lambda_{i+1}(Y)$. Then, each eigenvalue of $X$ can be bounded by, % \begin{equation} \lambda_i(Y) \leq \lambda_i(X) \leq \lambda_{n-m+i}(Y), \quad i = 1,\ldots,m \end{equation} \end{thm} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % THEOREM: GERSCHGORIN DISC % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Also helpful is the Gerschgorin disc theorem \cite{golub2012matrix} as it provides a way to compute an upper bound to the largest eigenvalue of $\bar{W}$ for any given choice of the driver nodes. % \begin{thm}[Gerschgorin Disc Theorem \cite{golub2012matrix}] Let $A = \{A_{i,j}\}$ be an $n \times n$ square matrix. The eigenvalues of $A$, $\lambda(A)$, lie in the union of the $n$ discs $D_i$, $i = 1, \ldots, n$ in the complex plane, each centered at $A_{i,i}$ with radius $C_i = \sum_{j=1,j\neq i}^n A_{i,j}$. The largest eigenvalue of the matrix $A$ can thus be upper bounded by, % \begin{equation} \lambda_{\max}(A) \leq \max\limits_{1<i<n} \sum_{j=1}^n A_{i,j} \end{equation} \end{thm} % \begin{cor} It is not difficult to show from Eqs. \eqref{eq:Wnn} and \eqref{eq:init} that $G_{0,0} > G_{i',j'}$ for $|i'|+|j'| > 0$. Then, the largest eigenvalue of $\bar{W}$ for any choice of the driver node, $v_a$, can be upper bounded by, % \begin{equation} \lambda_{\max}(\bar{W}) \leq \max\limits_{v_i \in \mathcal{T}} \sum_{v_j \in \mathcal{T}} G_{i',j'} \leq n_t G_{0,0} \end{equation} % where we use the shift $i' = i-a$ and $j'=j-a$. \end{cor} % \begin{figure} \centering \includegraphics[width=\columnwidth]{scaling.pdf} \caption{ The lower bound of the control metrics as a function of index of node further from the driver node. The NNOD lattice has edge weights $p = 3$ and $s = 1$ and driver node $\mathcal{D} = \{v_0\}$. The set of target nodes $\mathcal{T} \subset \{v_0,v_1,\ldots,v_7\}$. (A) The trace of the inverse of the output controllability Gramian for all sets of target nodes with cardinality $|\mathcal{T}|=n_t = 1,2,3$. The dashed line represents Eq. \eqref{eq:trinv}. (B) the negative log determinant of the output controllability Gramian for all sets of target nodes with cardinality $|\mathcal{T}| = n_t = 1,2,3$. The dashed line represents Eq. \eqref{eq:logdet}. } \label{fig:scale} \end{figure} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % SUBSUBSECTION: INVERSE OF TRACE % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsubsection{Inverse of the Trace} % Let $\mathcal{B} = \left\{\left. \textbf{y} \in \mathcal{R}^{n_t} \right|||\textbf{y}|| = 1 \right\}$ be the $n_t$-dimensional ball centered at the origin. The average minimum energy required to reach a point $\textbf{b}$, located on the unit hypersphere, from the origin, is found by integrating over the $\mathcal{B}$ \cite{summers2014optimal,summers2016submodularity}. % \begin{equation} \frac{\int_{\mathcal{B}} \textbf{y}^T \bar{W}^{-1} \textbf{y} d\textbf{y}}{\int_{\mathcal{B}} d\textbf{y}} = \frac{1}{n_t} \text{Tr} (\bar{W}^{-1}) \end{equation} %The minimum control energy, Eq. \eqref{eq:optcost}, can be expanded in the eigendecomposition $\bar{W} \boldsymbol{\xi}_i = \lambda_i \boldsymbol{\xi}_i$, % %\begin{equation} % J^* = \sum_{i=1}^{n_t} \left( \textbf{b}^T \boldsymbol{\xi}_i \right)^2 \frac{1}{\lambda_i} %\end{equation} % %The minimum energy is a weighted sum of the inverse eigenvalues of $\bar{W}$. %The trace of $\bar{W}^{-1}$ provides an indication of the order of the minimum energy \cite{summers2014optimal,summers2016submodularity}. %so that the trace of $\bar{W}^{-1}$ attempts to capture the order of the minimum energy for a generic choice of $\textbf{b}$. % Let $n_t > 1$ and $v_{\ell'} \in \mathcal{T}$ such that $\ell' = \max\limits_{v_{i'} \in \mathcal{T}} i'$, i.e., the index of the node furthest from the single driver node. Using the Cauchy interlacing theorem which states that $\lambda_1(\bar{W}) < G_{\ell',\ell'}$, we can lower bound the trace of the inverse of the output controllability Gramian. % \begin{equation}\label{eq:trinv} \begin{aligned} \text{Tr}(\bar{W}^{-1}) &\geq \frac{1}{\lambda_1(\bar{W})} \geq \frac{1}{G_{\ell',\ell'}} \sim \tilde{z}^{-\ell'} \end{aligned} \end{equation} % From the above expression we see that, as $0 < \tilde{z} < 1$, by reducing $\ell'$, we can reduce the lower bound. An example of the lower bound in Eq. \eqref{eq:trinv} is shown in Fig. \ref{fig:scale}(A) for a NNOD lattice with edge weights $p = 3$ and $s = 1$. Namely, the figure shows that by reducing $\ell'$ one can exponentially decrease the trace of the inverse output Controllability Gramian. This indicates that a convenient choice of an alternate driver node is the one that reduces the distance to the furthest target node. % \subsubsection{Volume of the Controllability Ellipsoid} % The controllability ellipsoid is defined as the set of vectors $\textbf{y}$ such that, % \begin{equation} \mathcal{S} = \left\{\textbf{y} \in \mathcal{R}^{n_t} | \textbf{y}^T \bar{W}^{-1} \textbf{y} \leq 1\right\}\end{equation} % The volume of $\mathcal{S}$ \cite{summers2016submodularity} is % \begin{equation}\label{eq:volume} V(\bar{W}) = \frac{\pi^{n_t/2}}{\Gamma(n_t/2+1)} \left(\det \bar{W}\right)^{1/n_t} \end{equation} % Note that $cV(\bar{W}) = V(c\bar{W})$, where, from Gerschgorin theorem we see that by setting $c = \frac{1}{n_t G_{0,0}}$, $\lambda_{n_t} (c\bar{W}) < 1$. % Minimizing $-\log \det (c\bar{W})$ can be interpreted as maximizing the volume of the reachable subspace for any value $J^*$. % \begin{equation}\label{eq:logdet} \begin{aligned} -\log\det (c\bar{W}) &\geq -\log (\lambda_1 (c\bar{W}))\\ &\geq - \log (cG_{\ell',\ell'}) \sim - \log (\tilde{z}) \ell' \end{aligned} \end{equation} % As $0 < \tilde{z} < 1$, $-\log(\tilde{z}) > 0$, and so the lower bound of -$\log \deg c (\bar{W})$ scales with $\ell'$. An example of the lower bound in Eq. \eqref{eq:logdet} is shown in Fig. \ref{fig:scale}(B). Namely, the figure shows that the bound holds tightly for $n_t = 1$ but becomes more conservative as the number of targets increase. By selecting a driver node such that $\ell'$ is minimized the $\log \det (\bar{W})$ may exponentially decrease the minimum control energy. %) %\begin{cor} % Another set of quantities of interest are the axis values, $G_{i',0}$. % Writing the Lyapunov equation from Eq. \eqref{eq:Wdot} for this value yields the relation, % % \begin{equation}\label{eq:axis} % \frac{2p}{s} G_{i',0} + G_{i'+1,0} + G_{i'-1,0} + 2G_{i',1} = 0 % \end{equation} % % where we use the symmetry $G_{i',-1} = G_{i',1}$. % Define $y_{i'} = G_{i',0} / G_{i'-1,0}$ for $i' > 0$ and rewrite Eq. \eqref{eq:axis} in terms of $y_{i'}$, % % \begin{equation} % y_{i'+1} y_{i'} = -\left(\frac{2p}{s} + 2f(i')\right) y_{i'} - 1 % \end{equation} % % where $f(i') = G_{i',1} / G_{i',0}$. % As $0 < G_{i',1} < G_{i',0}$, $0 < f(i') < 1$. %\end{cor} % \section{Discussion and Conclusion} % We have derived the exact expressions of the controllability Gramian for networked systems where the underlying topology is a lattice. We have specialized the results to NNOD lattices so that the length of the shortest path between two nodes appears in the expression for the Gramian entries. %While we derive analytical expressions for the NNOD lattice, exponential scaling of control energy metrics with respect to distance metrics has been reported for finite graphs of general topology \cite{chen2016energy,wang2017physical}. Our analytical expressions are in agreement with previously reported observations \cite{chen2016energy,wang2017physical}, that control energy metrics scale exponentially with respect to distance metrics for finite graphs with general topology. % Current research into driver node selection methods %either (i) to satisfy constraints placed on control energy metrics while minimizing the number of driver nodes or (ii) to minimize a control energy metric with respect to a fixed number of driver nodes, require the computation of many controllability Gramians (and their log determinant or inverse trace) which becomes prohibitively expensive for large networks in terms of both computation and storage. % Instead, the positive correlation between distance metrics with respect to the driver nodes and target nodes and the energy metrics discussed suggests that one may construct a heuristic driver node selection method based purely on the topology of the underlying graph. % %Methods from discrete location theory such as the facility location problem or $k$-median problem \cite{mirchandani1990discrete} to place driver nodes have the potential to be much more computationally efficient. %From initial comparative analysis to current Gramian based methods \cite{summers2016submodularity,tzoumas2016minimal}, discrete location based methods yield results comparative in terms of the final cost (either negative log determinant or trace of the inverse). An exhaustive numerical comparison between the greedy approximation algorithms \cite{summers2016submodularity,tzoumas2016minimal} and heuristic algorithms that minimize a distance metric between the sets of driver nodes and target nodes is forthcoming in a future publication. % \begin{ack} This work is supported by the National Science Foundation through NSF grant CMMI-1400193, NSF grant CRISP-1541148, and ONR Award No. N00014-16-1-2637 as well as HDTRA1-13-1-0020. \end{ack} % \bibliographystyle{plain} \begin{thebibliography}{10} \bibitem{benner2013numerical} Peter Benner and Jens Saak. \newblock Numerical solution of large and sparse continuous time algebraic matrix riccati and lyapunov equations: a state of the art survey. \newblock{\em GAMM-Mitteilungen}, 36(1):32--52, 2013. \bibitem{chen2016energy} Yu-Zhong Chen, Le-Zhi Wang, Wen-Xu Wang, and Ying-Cheng Lai. \newblock Energy scaling and reduction in controlling complex networks. \newblock{\em Royal Society open science}, 3(4):160064, 2016. \bibitem{cowan2012nodal} Noah~J. Cowan, Erick~J. Chastain, Daril~A. Vilhena, James~S. Freudenberg, and Carl~T. Bergstrom. \newblock Nodal dynamics, not degree distributions, determine the structural controllability of complex networks. \newblock{\em PloS one}, 7(6):e38398, 2012. \bibitem{gao2014target} Jianxi Gao, Yang-Yu Liu, Raissa~M. D'souza, and Albert-L{\'a}szl{\'o} Barab{\'a}si. \newblock Target control of complex networks. \newblock{\em Nature communications}, 5:5415, 2014. \bibitem{golub2012matrix} Gene~H. Golub and Charles~F. Van~Loan. \newblock{\em Matrix computations}, volume~3. \newblock JHU Press, 2012. \bibitem{katsura1971lattice} Shigetoshi Katsura, Tohru Morita, Sakari Inawashiro, Tsuyoshi Horiguchi, and Yoshihiko Abe. \newblock Lattice green's function. introduction. \newblock{\em Journal of Mathematical Physics}, 12(5):892--895, 1971. \bibitem{klickstein2017energy} Isaac Klickstein, Afroza Shirin, and Francesco Sorrentino. \newblock Energy scaling of targeted optimal control of complex networks. \newblock{\em Nature communications}, 8:15145, 2017. \bibitem{lin1974structural} Ching-Tai Lin. \newblock Structural controllability. \newblock{\em IEEE Transactions on Automatic Control}, 19(3):201--208, 1974. \bibitem{lin2011augmented} Fu~Lin, Makan Fardad, and Mihailo~R. Jovanovic. \newblock Augmented lagrangian approach to design of structured optimal state feedback gains. \newblock{\em IEEE Transactions on Automatic Control}, 56(12):2923--2929, 2011. \bibitem{liu2016control} Yang-Yu Liu and Albert-L{\'a}szl{\'o} Barab{\'a}si. \newblock Control principles of complex systems. \newblock{\em Reviews of Modern Physics}, 88(3):035006, 2016. \bibitem{liu2011controllability} Yang-Yu Liu, Jean-Jacques Slotine, and Albert-L{\'a}szl{\'o} Barab{\'a}si. \newblock Controllability of complex networks. \newblock{\em Nature}, 473(7346):167, 2011. \bibitem{mirchandani1990discrete} Pitu~B. Mirchandani and Richard~L. Francis. \newblock{\em Discrete location theory}. \newblock John Wiley \& Sons, Inc, 1990.\bibitem{morita1971useful} Tohru Morita. \newblock Useful procedure for computing the lattice green's function-square, tetragonal, and bcc lattices. \newblock{\em Journal of mathematical physics}, 12(8):1744--1747, 1971. \bibitem{olshevsky2014minimal} Alex Olshevsky. \newblock Minimal controllability problems. \newblock{\em IEEE Transactions on Control of Network Systems}, 1(3):249--258, 2014. \bibitem{pequito2017robust} S{\'e}rgio Pequito, Guilherme Ramos, Soummya Kar, A.~Pedro Aguiar, and Jaime Ramos. \newblock The robust minimal controllability problem. \newblock{\em Automatica}, 82:261--268, 2017. \bibitem{summers2016actuator} Tyler Summers. \newblock Actuator placement in networks using optimal control performance metrics. \newblock In {\em Decision and Control (CDC), 2016 IEEE 55th Conference on}, pages 2703--2708. IEEE, 2016. \bibitem{summers2016submodularity} Tyler~H. Summers, Fabrizio~L. Cortesi, and John Lygeros. \newblock On submodularity and controllability in complex dynamical networks. \newblock{\em IEEE Transactions on Control of Network Systems}, 3(1):91--101, 2016. \bibitem{summers2014optimal} Tyler~H. Summers and John Lygeros. \newblock Optimal sensor and actuator placement in complex dynamical networks. \newblock{\em IFAC Proceedings Volumes}, 47(3):3784--3789, 2014. \bibitem{sun2013controllability} Jie Sun and Adilson~E. Motter. \newblock Controllability transition and nonlocality in network control. \newblock{\em Physical review letters}, 110(20):208701, 2013. \bibitem{tang2014synchronization} Yang Tang, Feng Qian, Huijun Gao, and J{\"u}rgen Kurths. \newblock Synchronization in complex networks and its application--a survey of recent advances and challenges. \newblock{\em Annual Reviews in Control}, 38(2):184--198, 2014. \bibitem{tzoumas2015minimal} Vasileios Tzoumas, Mohammad~Amin Rahimian, George~J. Pappas, and Ali Jadbabaie. \newblock Minimal actuator placement with optimal control constraints. \newblock In {\em American Control Conference (ACC), 2015}, pages 2081--2086. IEEE, 2015. \bibitem{tzoumas2016minimal} Vasileios Tzoumas, Mohammad~Amin Rahimian, George~J. Pappas, and Ali Jadbabaie. \newblock Minimal actuator placement with bounds on control effort. \newblock{\em IEEE Transactions on Control of Network Systems}, 3(1):67--78, 2016. \bibitem{wang2015passivity} Jin-Liang Wang, Huai-Ning Wu, and Tingwen Huang. \newblock Passivity-based synchronization of a class of complex dynamical networks with time-varying delay. \newblock{\em Automatica}, 56:105--112, 2015. \bibitem{wang2017physical} Le-Zhi Wang, Yu-Zhong Chen, Wen-Xu Wang, and Ying-Cheng Lai. \newblock Physical controllability of complex networks. \newblock{\em Scientific reports}, 7:40198, 2017. \bibitem{yan2012controlling} Gang Yan, Jie Ren, Ying-Cheng Lai, Choy-Heng Lai, and Baowen Li. \newblock Controlling complex networks: How much energy is needed? \newblock{\em Physical review letters}, 108(21):218703, 2012. \bibitem{yan2015spectrum} Gang Yan, Georgios Tsekenis, Baruch Barzel, Jean-Jacques Slotine, Yang-Yu Liu, and Albert-L{\'a}szl{\'o} Barab{\'a}si. \newblock Spectrum of controlling and observing complex networks. \newblock{\em Nature Physics}, 11(9):779, 2015. \bibitem{yu2010some} Wenwu Yu, Guanrong Chen, and Ming Cao. \newblock Some necessary and sufficient conditions for second-order consensus in multi-agent dynamical systems. \newblock{\em Automatica}, 46(6):1089--1095, 2010. \bibitem{yu2009pinning} Wenwu Yu, Guanrong Chen, and Jinhu L{\"u}. \newblock On pinning synchronization of complex dynamical networks. \newblock{\em Automatica}, 45(2):429--435, 2009. \bibitem{yuan2013exact} Zhengzhong Yuan, Chen Zhao, Zengru Di, Wen-Xu Wang, and Ying-Cheng Lai. \newblock Exact controllability of complex networks. \newblock{\em Nature communications}, 4:2447, 2013. \bibitem{zhang2017efficient} Xizhe Zhang, Huaizhen Wang, and Tianyang Lv. \newblock Efficient target control of complex networks based on preferential matching. \newblock{\em PloS one}, 12(4):e0175375, 2017. \bibitem{zhao2005onset} Liang Zhao, Ying-Cheng Lai, Kwangho Park, and Nong Ye. \newblock Onset of traffic congestion in complex networks. \newblock{\em Physical Review E}, 71(2):026125, 2005. \bibitem{zhou2015controllability} Tong Zhou. \newblock On the controllability and observability of networked dynamic systems. \newblock{\em Automatica}, 52:63--75, 2015. \end{thebibliography} % \end{document} }
\caption{\textcolor{blue}{Maximal} Jacobian-based Saliency Map Attack (\textcolor{blue}{M-}JSMA\_F)}
\caption{Zurich Summer dataset: results on the five test images. (a) original image; (b) ground truth; (c) CNN, pixel-based; %(d) 2L$\lightning$CRF, pixel level; (d) CNN, region-level; (e) 2L$\lightning$CRF, region-level. For the accuracies of the single maps, please refer to Tab.~\ref{tab:num} (color legend: \textcolor{resi}{residential}, \textcolor{street}{street}, \textcolor{tr}{trees}, \textcolor{mea}{meadows}, \textcolor{rail}{railway}, \textcolor{wat}{water}, \textcolor{pool}{swimming pools}, \textcolor{bare}{bare soil}). \label{fig:resMaps}}
\caption{Zeebruges dataset (\texttt{grss\_dfc\_2015}): results on the two test tiles. (a) original image; (b) ground truth; (c) CNN, pixel-based; (d) 2L$\lightning$CRF, pixel level. For the accuracies of the single maps, please refer to Tab.~\ref{tab:numZ} (color legend: \textcolor{Zimp}{impervious (white colored)}, \textcolor{Zwater}{water}, \textcolor{Zclutter}{clutter}, \textcolor{ZlowVeg}{low vegetation}, \textcolor{Zbuilding}{buildings}, \textcolor{Ztree}{trees}, \textcolor{Zboat}{boats}, \textcolor{Zcar}{cars}). The ground truth images are blurred as in~\cite{DFCA}, since they are undisclosed. \textbf{(Note: in this preprint we had to decrease the graphics resolution. For full resolution, please refer to the published version, or contact the authors)}.\label{fig:resMapsZ}}
\caption{Zeebruges dataset (\texttt{grss\_dfc\_2015}): zoomed results on details of the two test tiles of Fig.~\ref{fig:resMapsZ}. (a) original image; (b) ground truth; (c) CNN, pixel-based; (d) 2L$\lightning$CRF, pixel level. For the accuracies of the single maps, please refer to Tab.~\ref{tab:numZ} (color legend: \textcolor{Zimp}{impervious (white coloured)}, \textcolor{Zwater}{water}, \textcolor{Zclutter}{clutter}, \textcolor{ZlowVeg}{low vegetation}, \textcolor{Zbuilding}{buildings}, \textcolor{Ztree}{trees}, \textcolor{Zboat}{boats}, \textcolor{Zcar}{cars}). The ground truth images are blurred as in~\cite{DFCA}, since they are undisclosed. \textbf{(Note: in this preprint we had to decrease the graphics resolution. For full resolution, please refer to the published version, or contact the authors)}. \label{fig:resMapsZZ}}
\caption{\linespread{1.2}\small Axial (a) and sagittal (f) views of a 3DCT scan, (b,g) attention coefficients, \bluereview{image feature activations before (c,h) and after attention gating (d,e,i,j)}. Similarly, (k-n) visualise the gating on a coarse scale skip connection. The filtered feature activations (d,e,i,j) are collected from multiple AGs, where a subset of organs is selected by each gate and activations consistently correspond to specific structures across different scans.}
\caption{ Asynchronous RAS with additive coarse grid in local form. Variables printed in {\color{blue}blue} are exposed memory regions that are local to the calling process. {\color{red}Red} variables are remote memory regions. }
\caption{Translations generated by VAG-NMT and Text-Only NMT. VAG-NMT performs better in the first two examples, while Text-Only NMT performs better in the third example. We highlight the words that distinguish the two systems' results in \textcolor{red}{red} and \textcolor{blue}{blue}. \mingyang{\textcolor{red}{Red} words are marked for better translation from VAG-NMT and \textcolor{blue}{blue} words are marked for better translation from Text-Only NMT. }}
\caption{Average PSNR/SSIM values for x2 scale factor for various models on different models. \textcolor{red}{red} indicates best value and \textcolor{blue}{blue} indicates second best value.}
\caption{FLOPs count (x1e6) for various models with suitable input to produce a size 64x64 output image and scale factor of 2. \textcolor{red}{red} indicates best value and \textcolor{blue}{blue} indicates second best value.}
\caption{ % Inconsistent results produced by \DAM{} on automatically generated adversarial examples. % The notation \sout{segment one} \textcolor{red}{segment two} denotes that the corruption process removes ``segment one'' and introduced ``segment two'' in the sentence, and $s_{1} \textcolor{red}{\xrightarrow{p}} s_{2}$ indicates that \DAM{} classifies the relation between $s_{1}$ and $s_{2}$ as \emph{contradiction}, with probability $p$. % We use different colours for representing the \textcolor{red}{contradiction}, \textcolor{green}{entailment} and \textcolor{blue}{neutral} classes. % Examples 1, 2, 3, and 4 violate the rule $\xrule{2}$, while example 5 violates the rule $\xrule{5}$. $.00 \rightsquigarrow .99$ indicates that the corruption process increases the inconsistency loss from .00 to .99, and the \cfbox{red}{red boxes} are used for indicating mistakes made by the model on the adversarial examples. % }
\caption{ % Inconsistent results yield by \DAM{} on the SNLI training set. % The notation $s_{1} \textcolor{red}{\xrightarrow{p}} s_{2}$ indicates that \DAM{} classifies the relation between $s_{1}$ and $s_{2}$ as \emph{contradiction} with probability $p$. % We use different colours for representing the \textcolor{red}{contradiction}, \textcolor{green}{entailment} and \textcolor{blue}{neutral} classes. % Examples $\{ 1, 2, 3\}$ (resp. $\{ 4, 5, 6 \}$) violate the logic rule $\xrule{2}$ (resp. $\xrule{3}$) in \cref{tab:rules}. }
\caption{ % Inconsistent results produced by \DAM{} on adversarial examples generated using the discrete search procedure described in \cref{ssec:discrete} -- the pattern \sout{segment one} \textcolor{red}{segment two} denotes that the corruption process replaced ``segment one'' with ``segment two''. % Examples $\{ 1, 2, 3 \}$ (resp. $\{ 4, 5, 6 \}$) violate the rule $\xrule{2}$ (resp. $\xrule{4}$), while examples $\{ 7, 8, 9 \}$ violate the logic rule in $\xrule{5}$. % }
\caption{ The ratio ($\chi^2/\chi^2_0$) of total $\chi^2$ values (all data sets combined) as a function of the a)~charm and b)~bottom matching scale $\mu_{c,b}$ in GeV. $\chi_{0}^{2}$ is the $\chi^{2}$ value for $\mu_{m}$ equal to the quark mass. % The triangles (blue \textcolor{blue}{$\blacktriangle$} ) are NLO and the diamonds (red \textcolor{red}{$\blacklozenge$}) are NNLO. % The fits are from Ref.~\cite{Bertone:2017ehk}. \label{fig:chi2scaledi} % }
\caption{Performance on the Flickr30k Karpathy test split. The symbol \ssymbol{1} denotes directly optimizing CIDEr. The symbol \ssymbol{2} denotes using extra data for training, thus not directly comparable. Nonetheless, our model supersedes all existing models in SPICE, which correlates the best with human judgments.}
\caption{Performance on the COCO Karpathy test split. Symbols, \ssymbol{1} and \ssymbol{2}, are defined similarly. Our model outperforms the current state-of-the-art Up-Down substantially in terms of SPICE.\label{tab:res-mscoco}}
\caption{Performance on the online COCO evaluation server. The SPICE metric is unavailable for our model, thus not reported. c5 means evaluating against 5 references, and c40 means evaluating against 40 references. The symbol \ssymbol{1} denotes directly optimizing CIDEr. The symbol \ssymbol{2} denotes model ensemble. The symbol \ssymbol{3} denotes using extra data for training, thus not directly comparable. Our submission does not use the three aforementioned techniques. Nonetheless, our model is second only to Up-Down and surpasses almost all the other models in published work, especially when 40 references are considered.\label{tab:res-server}}
\caption{Classification accuracy and response distribution entropy for \textcolor{googlenet.100}{GoogLeNet}, \textcolor{vgg.100}{VGG-19} and \textcolor{resnet.100}{ResNet-152} as well as for \textcolor{human.100}{human observers}. `Entropy' indicates the Shannon entropy of the response/decision distribution (16 classes). It here is a measure of bias towards certain categories: using a test dataset that is balanced with respect to the number of images per category, responding equally frequently with all 16 categories elicits the maximum possible entropy of four bits. If a network or observer responds prefers some categories over others, entropy decreases (down to zero bits in the extreme case of responding with one particular category all the time, irrespective of the ground truth category). Human `error bars' indicate the full range of results across participants. Image manipulations are explained in Section~\ref{meth:image_manipulations} and visualised in Figures~\ref{fig:stimuli_noise_contrast}, \ref{fig:stimuli_lowpass_highpass}, \ref{fig:stimuli_eidolon_I_II}, \ref{fig:stimuli_eidolon_III_phase_scrambling_false_colour_power_equalisation} and \ref{fig:stimuli_salt_and_pepper_noise}.}
\caption{Classification accuracy (in percent) for networks with potentially distorted training data. Rows show different test conditions at an intermediate difficulty (exact condition indicated in brackets, units as in Figure~\ref{fig:results_accuracy_entropy}). Columns correspond to differently trained networks (leftmost column: human observers for comparison; no human data available for salt-and-pepper noise). All of the networks were trained from scratch on (a potentially manipulated version of) 16-class-ImageNet. Manipulations included in the training data are indicated by a \textcolor{human.100}{red} rectangle; additionally `greyscale' is underlined if it was part of the training data because a certain distortion encompasses greyscale images at full contrast. Models~\textbf{A1~to~A9}: ResNet-50 trained on a single distortion (100 epochs). Models~\textbf{B1~to~B9}: ResNet-50 trained on uniform noise plus one other distortion (200 epochs). Models~\textbf{C1~\&~C2}: ResNet-50 trained on all but one distortion (200 epochs). Chance performance is at $\frac{1}{16}=6.25\%$ accuracy.}
\caption{Mean entropy of the probabilities for the 1000 ILSVRC classes for \textcolor{googlenet.100}{GoogLeNet}, \textcolor{vgg.100}{VGG-19} and \textcolor{resnet.100}{ResNet-152}. Dotted line indicates the maximum possible entropy. This is a measure of network `uncertainty'.}
\caption{Classification accuracy and response distribution entropy for \textcolor{human.100}{human observers}, \textcolor{resnet50.100}{ResNet-50} as well as an \textcolor{all.distortions.net.100}{All-Distortions-Net} and a \textcolor{specialised.net.100}{Specialised-Net}. All networks are trained from scratch; the Specialised-Net in every plot is trained on a single distortion (models A1 to A9 in Figure~\ref{fig:results_training}) whereas the All-Distortions-Net is trained on a number of distortions simultaneously. This corresponds to models C1 and C2 in Figure~\ref{fig:results_training}: for subplot \ref{sup:training_salt_and_pepper_noise}, salt-and-pepper noise, performance of model C1 is shown. For subplot \ref{sup:training_uniform_noise}, uniform noise, performance of C2 is shown. For all other plots, performance of the All-Distortions-Net is shown as the mean of performance for models C1 and C2.}
\caption{Decay of the dark soliton at $\beta=0$. $|\psi_+|$ and $|\psi_-|$ components are shown in the left and right columns, respectively. For $t=27$ we also show phases $\textrm{arg}\,\psi_\pm$. All distributions are shown within $(x,y) \in[-15,15]\times [-33.4,33.4]$ windows. See detailed dynamics in \textcolor{blue}{Visualization 1}. \label{fig:ev30}}
\caption{Decay of the dark soliton at $\beta=-0.1$. $\psi_+$ and $\psi_-$ are shown in left and right columns, respectively, within $(x,y) \in [-15,15] \times [-33.4,33.4]$ windows. For $t=50$ both amplitudes and phases are shown. See \textcolor{blue}{Visualization 2}. \label{fig:ev22}}
\caption{Decay of dark soliton at $\beta=0.1$. $\psi_+$ and $\psi_-$ are shown in left and right columns, respectively, within $(x,y) \in [-15,15]\times [-33.4,33.4]$ windows. For $t=27$ both amplitudes and phases are shown. See \textcolor{blue}{Visualization 3}. \label{fig:ev23}}
\caption{Decay of the half-dark soliton with $\beta=0.4$. $\psi_+$ and $\psi_-$ correspond to left and right columns, respectively. For $t=35$ both amplitudes and phases are shown. All panels are shown within $(x,y) \in [-10,10]\times [-33.4,33.4]$ windows. See \textcolor{blue}{Visualization 4}. \label{fig:ev28}}
\caption{P-wave: $ \varepsilon_a^{\text{\tiny $\bigcirc$}} = 0.11\% $}
\caption{S-wave: $ \varepsilon_a^{\text{\tiny $\bigcirc$}} = 0.11\% $}
\caption{Quasienergy gap width of the bulk spectrum as a function of the relative shift between the sawtooth intensity profile and the optical lattice. All quasienergy gaps turn out to be identical. The graph is shown for the representative case of a magnetic flux $\phi=1/3$. The three curves refer to the case of a simple lattice (\textcolor[RGB]{94,129,181}{$\bullet$}), a superlattice with a two-fold longer lattice constant (\textcolor[RGB]{255,156,36}{\mbox{\tiny $ \blacksquare $}}), and a superlattice with a four-fold longer lattice constant (\textcolor[RGB]{143,176,50}{$\blacktriangle$}). The relative shift is expressed in units of the rescaled lattice constant. The quasienergy gap width is maximum at a relative shift of $0.5$, when the sawtooth intensity profile is aligned with respect to the lattice sites as shown in the illustration in Fig.~\ref{fig:SawtoothIntensityPattern}. }
\caption{Some ``gold" references and best hypotheses (after rescoring by different language models) for eval92 set. In red are errors or missing word (denoted as `\textcolor{red}{\textvisiblespace}').}
\caption{Algorithm 2 (Sampling procedure) of IMM as in the original paper~\cite{tang15}.}{\includegraphics[width=0.7\linewidth]{IMMSamplingAlgo.png}\label{fig:sampling}}
\caption{\label{fig:ce_error}Cross entropy error on dev set of NOHIER (\textcolor{red}{red}) and HAQAE (\textcolor{blue}{blue}) models as training progresses.}
\caption{Sample outputs from the baseline and our proposed system. The seeds (what is given to the system) are shown in the left column while the outputs are on the right. HAQAE is able to distinguish between the contrasting seeds. \textcolor{red}{Red} highlights the lack of branching quality in the baseline model and \textcolor{blue}{Blue} highlights the correct behavior as exhibited by HAQAE.}
\caption{\color{red}place holder for structured prediction. Reference to the noise preprocessing is here on the footnote \footnotemark[1]}
\caption{\color{red}place holder for structured prediction. Reference to the noise preprocessing is here on the footnote \footnotemark[1]}
\caption{\myred{This figure has been deleted.}}
\caption{ A mis-grouping occurs when the cake in \textcolor{red}{red} box is grouped with the child in \textcolor{green}{green} box to form a triplet $\langle human, cut, cake \rangle$. Such negative triplets have not been annotated with any HOI labels. We specifically mine for such hard negative triplets during training, to let the model learn to better discriminate mis-grouped triplets from the correct ones. }
\caption{ Detections with triplet score larger than 0.7 are displayed. Without the proposed negative triplet mining, the model gives all four predictions, in which three of them (\ie~with \textcolor{red}{red} headings) are mis-grouped. Model with negative triplet mining reduces the prediction to only the correct triplet (\ie~the bottom right with \textcolor{blue}{blue} heading). }
\caption{ The effect of human intention in V-COCO and HICO-DET. HOI predictions together with the triplet scores (with \textcolor{gray}{gray} headings) are shown. After leveraging intention ({\it VTransE w/ P+G} vs. {\it VTransE}), we show the change in triplet scores for the detections. Using intention suppresses the false prediction scores (\ie~column 1 -- 2 with \textcolor{red}{red} headings), whereas improves the correct ones (\ie~column 3 -- 5 with \textcolor{blue}{blue} headings). The corresponding gaze density heatmaps intuitively demonstrate that fixated regions are informative of HOIs. The pose information is not plotted. Triplet scores are obtained as in Section~\ref{subsubsec:inference}. }
\caption{(Color online) Rendering - to scale - of the terminal part of the low energy section. The borosilicate glass science chamber is shown on the right hand side. The ion beam enters from the CF40 connection on the left, next to the \SI{20}{\litre\per\second} ion pump, passes through the gate valve, and enters the laser cooling chamber. A CF16 blind flange equipped with electrical feedthroughs supports an implantation foil. Anti-Helmholtz quadrupole coils are visible at either side of the glass chamber. \red{The implantation thin foil (not visible) is placed \SI{20}{\milli\meter} inside the science chamber. Further details are shown in Fig.~\ref{fig:neutralizer}.} Positions of the laser beams are marked by red cylinders. A solid reservoir of $^{133}$Cs is sealed by the CF16 valve visible in the background.}
\caption{(Color online) Sketch - to scale - of the neutralizer mount. A thin Y foil, where the incoming ion beam is focused, is supported by copper contacts, in turn connected to electrical feedthroughs. Mechanical stability is ensured by ceramic rods, secured to a CF16 flange. Resistive heating allows one to control the temperature of the foil, up to \SI{e3}{\kelvin}. \red{Top: relative position of the Y foil and the MOT (approximately at the center of the laser beams overlapping region). Bottom: detailed view of the neutralizer mount. }}
\caption{\red{(Color online) Direct loading from neutralized} $^{133}$Cs$^{+}$ beam: N$_{\text{MOT}}$ before (blue disks) and after (red diamonds) exposure to \SI{e-10}{\ampere} $^{133}$Cs$^{+}$ beam, as a function of T$_{\text{Y}}$. Inset: MOT fluorescence measured at T$_{\text{Y}}=\SI{691}{\kelvin}$ and \SI{887}{\kelvin} (I$_{\text{DC}}=\SI{3.5}{\ampere}$ and \SI{5}{\ampere}). The vertical dashed lines mark the timestamps of magnetic field gradient on (t=-\SI{30}{\second}), ion beam on (t=\SI{0}{\second}), ion beam off (t=\SI{300}{\second}), magnetic field gradient off (t=\SI{400}{\second}). The horizontal dashed lines mark the detection threshold.}
\caption{\textbf{Raster plots for the HH neuronal network in three different dynamical regimes.} Raster plots of $10$ randomly selected neurons for each case are shown. A short bar indicates that the neuron with\textcolor{red}{{} }certain index fires at certain time. The coupling strength is selected at random from the uniform distribution of the interval $[0,s]$, where $s=\unit[0.071]{ms^{-1}}$ (the corresponding physiological excitatory postsynaptic potential is $\sim\unit[1]{mV}$). The Poisson input parameters for in (a), (b), and (c) are $(\mu=\unit[0.6]{ms^{-1}},\,f=\unit[0.05]{ms^{-1}})$, $(\mu=\unit[1.1]{ms^{-1}},\,f=\unit[0.04]{ms^{-1}})$ and $(\mu=\unit[2.5]{ms^{-1}},\,f=\unit[0.03]{ms^{-1}})$, respectively. \label{fig:Raster}}
\caption{\textbf{Electrophysiological verification for the second-order MEP analysis with short-time recordings.} Spike trains of randomly selected eight neurons are recorded from V1 in \emph{anesthetized} macaque monkeys (see Appendix \ref{sec:pvc8Exp}). (a) A short bar indicates that the neuron with\textcolor{red}{{} }certain index fires at certain time. (b) Mean absolute values of effective interactions from the first-order to the fourth-order are plotted, with the standard deviation indicated by the error bar, computed by the long recording of $\unit[3824]{s}$. (c) The frequency of each firing state from the distribution measured in the short recording of $\unit[191.2]{s}$ (magenta) and the distribution of the second order MEP analysis, $P_{2}$ (blue), is plotted against the frequency measured from the long recording of $\unit[3824]{s}$. \label{fig:MEPExp}}
\caption{Top-5 most similar input and output representations to two query words based on cosine similarity for an NMT trained without (NMT) or with \textit{weight tying} (NMT-\code{tied}) and our \textit{structure-aware} output layer (NMT-\code{joint}) on De-En ($|\mathcal{V}| \approx 32K$). Our model learns representations useful for encoding and generation which are more consistent to the dominant semantic and syntactic relations of the query such as verbs in past tense, adjectives and nouns (inconsistent words are marked in \textcolor{red}{red}).}
\caption{ Working principle of coherent perfect absorption (A) %Coherent perfect absorption. Incident radiation enters an absorber from two sides with a wavevector $q_\star$. Provided the complex amplitudes $a^L$ and $b^R$ are chosen properly, no radiation is transmitted or reflected. (B) Experimental realization with a BEC in an optical lattice. The matter waves enter a lossy lattice site from both sides. The losses are realized with an electron beam, which removes the atoms. Due to the interactions between the atoms, the two incoming waves are nonlinear, \red{while the absorption in the lossy site is linear.} }
\caption{Illustration of the generated attentional maps of the crowd flow in {\color{red}{periodic representation learning}} with ${m}$ set as 2. Every three columns form one group. In each group: i) on the first row, the first two images are the input periodic inflow/outflow maps and the last one is the ground truth inflow/outflow map of next time interval; ii) on the second row, the first two images are the attentional maps generated by our ACFM, while the last one is our predicted inflow/outflow map; iii) on the third row, the first two images are the residual maps between the input flow maps and the ground truth, while the last one is the residual map between our predicted flow map and the ground truth.}
\caption{Illustration of the generated attentional maps of the crowd flow in {\color{red}{sequential representation learning}} with ${n}$ set as 3. Every four columns form one group. In each group: i) on the first row, the first three images are the input sequential inflow/outflow maps and the last one is the ground truth inflow/outflow map of next time interval; ii) on the second row, the first three images are the attentional maps generated by our ACFM, while the last one is our predicted inflow/outflow map; iii) on the third row, the first three images are the residual maps between the input flow maps and the ground truth, while the last one is the residual map between our predicted flow map and the ground truth.}
\caption{Examples on WMT2018 validation data. The source and translated sentences, the reference sentences, the predictions of the CEQE without and with POS tags and baseline features are shown. Words predicted as OK are shown in {\color{green}green}, those predicted as BAD are shown in {\color{red}red}, the difference between the translated and reference sentences are shown in {\color{blue}blue}.\label{tab:example}}
\caption{\label{fig:4} \red{The raster plot for a network of $N_E = 20,000$ excitatory (green) and $N_I = 5,000$ inhibitory (red) QIF neurons is displayed in (a). In (c) and (d) the COs' frequencies, measured from the power spectrum $S(\nu)$ of the mean voltage $V(t)$ (shown in (b)), are reported as symbols versus the excitatory DC current $I_0^{e}$. The dashed lines are the theoretical mean-field predictions. The values of the parameters are reported in \cite{model}.}}
\caption{\label{tab:output} Example output from TEMPL (top) and NPC+CC (bottom). Text that accurately reflects a record in the associated box or line score is in {\color{blue}blue}, erroneous text is in {\color{red}red.}}
\caption{Example documents from the template-based system, WS-2017, the best system of Wiseman et al.~\shortcite{wiseman2017challenges}, and our Neural Content Planning model with conditional copy (NCP+CC). Text that accurately reflects a record in the associated box or line score is recorded in {\color{blue}blue}, erroneous text is marked in {\color{red}red}, duplicate text is marked in {\color{orange}orange}.}
\caption{Ordinal Logistic Regression Models for the YouTube Experiment. Each Column Represents an Ordinal Regression Model. The Name of the Column Indicates the Dependent Variable. Values Within Brackets Represent the 95\% Confidence Interval.\\ \colorbox[rgb]{0.87,0.92,0.97}{$\ast$} $p < 0.05\enspace$ \colorbox[rgb]{0.78,0.86,0.94}{$\ast\ast$} $p < 0.01\enspace$ \colorbox[rgb]{0.62,0.79,0.88}{$\ast\ast\ast$} $p < 0.001\enspace$ \colorbox[rgb]{0.42,0.68,0.84}{$\ast\ast\ast\ast$} $p < 0.0001\enspace$}
\caption{Ordinal Logistic Regression Models for the Pinterest Experiment. Each Column Represents an Ordinal Regression Model. The Name of the Column Indicates the Dependent Variable. Values Within Brackets Represent the 95\% Confidence Interval.\\ \colorbox[rgb]{0.87,0.92,0.97}{$\ast$} $p < 0.05\enspace$ \colorbox[rgb]{0.78,0.86,0.94}{$\ast\ast$} $p < 0.01\enspace$ \colorbox[rgb]{0.62,0.79,0.88}{$\ast\ast\ast$} $p < 0.001\enspace$ \colorbox[rgb]{0.42,0.68,0.84}{$\ast\ast\ast\ast$} $p < 0.0001\enspace$}
\caption{The pipeline of our framework. E denotes the encoder, G denotes the generator, and D is the discriminator. S is the pixel-wise classifier for semantic segmentation. The red color represents the network blocks for the source domain, and the blue for the target domain. We also display the Conservative Loss and its backpropagation. \textcolor{red}{$\nearrow$} represents the gradient ascend and \textcolor{red}{$\searrow$} denotes the gradient descend}
\caption{\blue{Accuracy of superpixel segmentation (F1 score) for each dataset type when using different proportions of positive pixels within a superpixel to define a positive superpixel. F1 scores are averaged over 4 sequences.}}
\caption[\color{red}Caption for LOF]{Absolute position and peculiar motion of \fgl within the Galaxy in a view perpendicular to the Galactic plane ({\em top}) and parallel to it in the projected direction between the Galactic center and \fgl ({\em bottom}). The black dot represents the Galactic center, the orange star represents the Sun, the open red symbol represents SNR~G384.3$-$1.8, and the blue star represents \fgl at the distance of 6.4~kpc. The blue arrow represents the peculiar motion of the source (in the RSR, after Galactic rotation subtraction) after 70~Myr for reference\protect\footnotemark. The blue ellipse represents the uncertainty on this motion. The gray ellipses provide a schematic picture of the Galactic plane. {\em Image Credit:} NASA/JPL-Caltech/R.\Hurt/SSC/Caltech, The Mozilla Foundation.}
\caption{Two experimental snapshot MFM raw images. (a) Experiment 1: snapshot captured 2D MFM image of multiple static periplasms by using an MFG with $9$ focal planes under exposure time of $0.5$s. (b) Experiment 2: a frame from an MFM video of a moving bacterium captured at $25$ fps by using an MFG with $25$ focal planes. The raw MFM video is shown in \textcolor{urlblue}{Visualization 1}.}
\caption{Experimental 3D reconstructions of a movable bacterium. A raw MFM video (shown in \textcolor{urlblue}{Visualization 1}) was captured at $25$fps as the bacterial moves in 3D space. The computational 3D reconstruction was performed for each video frame. Five out of sixty frames reconstruction is shown in (a-e). (f) 3D trajectory of the bacterium by computing and tracking its center of mass for each frame reconstruction. The colorbar indicate the frame index over time. The complete 3D video reconstruction from the first frame to the last frame is shown in \textcolor{urlblue}{Visualization 2}.}
\caption{\blue{(a)} Illustration of the initial configuration of the system. The top of the box corresponds to the $^4$He surface, and a particle of radius $5$ $\mu$m (a red ball) is trapped at the top of a straight vertical vortex filament (a blue line). We excite the vertical filament by letting a small vortex ring (a blue ring) collide with it. \blue{(b)} The \blue{trapped} particle \blue{at the surface} after the excitation. Only the tip of the vortex bends and starts to oscillate.}
\caption{A typical particle trajectory on the surface. At the beginning it shows a large scale motion, but it decays quickly as a small scale circular motion of period \blue{$T_{traj}$} and radius \blue{$R_{traj}$} appears at around $0.6$ \blue{s} for this plot. }
\caption{(a) Period and (b) radius of the circular motion as a function of the particle density $\rho_p$. Both \red{$T_{traj}$} and \red{$R_{traj}$} values are obtained by averaging over some time after the spiral motion decays sufficiently\red{, and error bars are obtained by taking the standard deviations}. Since there are \blue{no} noticeable differences in the time-averaged motion in $x$ and $y$ -directions, each plot is represented by the motion in $x$-direction. Here, the particle \blue{radius $a_p$} is fixed to be $5$ $\mu$m. Both plots are fitted with a function of form $y=\alpha x^\beta + \gamma$ \blue{as shown in the legends} and represented by the black dashed curves. $\rho_p = 3594.7$ kg/m$^3$ corresponds to the density of Ba at room temperature.}
\caption{(a) Period \red{$T_{traj}$} and (b) radius \red{$R_{traj}$} of the circular motion as a function of the particle radius $a_p$. The \red{$T_{traj}$} and \red{$R_{traj}$} values are obtained by time-averaging in the same way as in Fig.\ref{fig:a-dependence}, and the plots are represented by the motions in $x$-direction. The particle density is set to be that of \blue{barium}. The plots are fitted with \blue{the functions shown in the legends}. }
\caption{(a) Illustration of the local radial counter flow around a heated particle. Normal flow is radiated away from the particle, while \red{superflow} is toward the particle. The figure (b) shows how a particle of radius $5$ $\mu$m attracts a vortex that is initially separated by a distance $50$ $\mu$m and reconnects with it in a cubic system of length $200$ $\mu$m. \blue{Filament positions are shown with a time step of} $4\times10^{-3}$ s, and the whole process occurs within $0.04$ s.}
\caption{{\bf Example of a Multimodal KB.} Graph representation of (a part of) a KB that consists of regular links (in black) and multimodal ones (in \textcolor{violet}{purple}) that we support in this work. }
\caption{The curricula for learning nested for loops in Code.org. To provide intuition on the vast domain complexity, we show the \textcolor{teal}{number of unique solutions} and the \textcolor{gold}{number of students} who attempted the problem for each of the 8 exercises.}
\caption{Mean AUC values for augmentation scenarios. Each color and marker represent a prediction method: \textcolor{color:circle}{$\bullet$} original image; \textcolor{color:triangle}{$\blacktriangle$} test-time data augmentation (64 images); \textcolor{color:square}{$\blacksquare$} 144 crops. Error bars represent the standard deviation for 6 runs. Values reported on ISIC Challenge 2017 test set.\vspace{-0.2cm}}
\caption{Mean AUC values for different training dataset sizes, randomly sampled from the ISIC Challenge 2017 training dataset. Colors and markers represent the use of data augmentation: \textcolor{color:square2}{$\blacksquare$} no data augmentation; \textcolor{color:triangle2}{$\blacktriangle$} train data augmentation (scenario J); \textcolor{color:circle2}{$\bullet$} train and test data augmentation (scenario J, averaging each test image on 64 augmented samples). Bands represent the standard deviation for 6~runs. Values reported on ISIC Challenge 2017 test set.\vspace{-0.2cm}}
\caption{The discriminator's performance can be interpreted through the \textit{layered distributions} view, a composite visualization composed of 3 layers selected by the user: \textit{Real samples}, \textit{Fake samples}, and \textit{Discriminator's classification}. Here, the discriminator is performing well, since most \textcolor{my_green}{real samples} lies on its classification surface's green region (and \textcolor{my_purple}{fake samples} on purple region). }
\caption{Evaluating how well the distribution of fake samples matches that of real samples by turning on \textcolor{my_green}{real samples' density contour} and \textcolor{my_purple}{fake samples} in the layered distributions view.}
\caption{Example of understanding the interplay between discriminator and generator using the layered distributions view. \textcolor{my_purple}{Fake samples}' movement directions are indicated by the generator's \textit{gradients} (pink lines), based on those samples' current locations and the discriminator's current classification surface (visualized by background colors). }
\caption{With \name{}, users can interactively train Generative Adversarial Networks (GANs), and visually examine the model training process. In this example, a user has successfully used \name{} to train a GAN that generates 2D data points whose challenging distribution resembles a ring. \textbf{A.} The \textit{model overview graph} summarizes a GAN model's structure as a graph, with nodes representing the \textcolor{my_purple}{generator} and \textcolor{my_blue}{discriminator} submodels, and the data that flow through the graph (e.g., fake samples produced by the generator). \textbf{B.} The \textit{layered distributions} view helps users interpret the interplay between submodels through user-selected layers, such as the discriminator's classification heatmap, \textcolor{my_green}{real samples}, and \textcolor{my_purple}{fake samples} produced by the generator. }
\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{\label{tab:personachoice} Retrieval precision on the \reddit{} test set using a \attentionmodel{} and different persona selection systems. $N$: maximum number of sentences per persona.}
\caption{\label{tab:transfer} hits@1 results for the best found \attentionmodel{} architecture on different test sets. FT-PC: \reddit{}-trained model fine-tuned on the \personachat{} training set. To be comparable to the state of the art on each dataset, results on \personachat{} are computed using 20 candidates, while results on \reddit{} use 100.}
\caption{Results for LAS (+ mono-treebank baseline), MLAS, sentence and word segmentation, UPOS tagging and morphological features (UFEATS). Treebanks sharing a parsing model grouped together; substitute and proxy treebanks for segmentation, tagging, parsing far right (\textsc{special} models detailed in the text). Confidence intervals for coloring: \colorbox{red!40}{$|$} $< \mu \!-\! \sigma <$ \colorbox{red!20}{$|$} $< \mu \!-\! %\sigma^- \textsc{se} < \mu < \mu \!+\! %\sigma^- \textsc{se} <$ \colorbox{green!20}{$|$} $< \mu \!+\! \sigma <$ \colorbox{green!50}{$|$}. }
\caption{The design process of the material from initial guess to final macro-structure represented in the unit cell and as a periodic material. \protect \newboxsymbol{magenta}{magenta} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{The design process of the material from initial guess to final macro-structure represented in the unit cell and as a periodic material. \protect \newboxsymbol{magenta}{magenta} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{The design process of the material from initial guess to final macro-structure. \protect \newboxsymbol{magenta}{magenta} Young modulus of $0.91$, \protect \newboxsymbol{yellow}{yellow} void.}
\caption{\red{Effect of common and separated grounds on a 2.4~GHz signal transmitted on a spraypainted sheet.}}
\caption[caption]{\red{{\bf Sharing the surface.} \textmd{Both the clients are in contact with the surface and use carrier sense to share the conductive surface. The first plot is when the two clients are on channel 1 and 6 respectively and the second plot is when they share the same Wi-Fi channel 1.}}}
\caption{\red{RSSI over distance. \textmd{}}}
\caption{\red{2x2 MIMO}}
\caption{\red{3x3 MIMO}}
\caption{\red{Condition number of Surface MIMO. \textmd{}}}
\caption{\red{2x2 MIMO}}
\caption{\red{3x3 MIMO}}
\caption{\red{{\bf Gbps surface communication. \textmd{The wideband nature of conductive surfaces can be leveraged to support Gbps communication. (a) A virtual reality headset streams video through a hoodie's conductive cloth to a wearable pack. (b) HD cameras on walls coated with conductive paint stream video to a plugged-in hub through the wall.}}}}
\caption[caption]{{\bf Signal attenuation over distance \red{for conductive spraypaint and cloth.}} \textmd{In the absence of contact \red{ with the surface, the received power} is close to \red{the noise floor}, demonstrating that communication between the devices occur due to their contact with the conductive surface.}}
\caption{Visual results on two videos for the object tracking, using maximum correlation of the detector to dynamically vary MR. For each video, the top row shows the frames acquired with no compression (conventional imaging, referred to as the oracle). The second and third rows display the reconstructions using Rate-Adaptive ReconNet (trained for MR = $[0.04, 0.10]$ and vanilla ReconNet (trained for MR = 0.10) respectively. \textcolor{blue}{Blue}, \textcolor{green}{green} and \textcolor{red}{red} boxes show the object locations for ground-truth, Rate-Adaptive ReconNet and vanilla ReconNet respectively. Unlike the rate-adaptive framework, as MR is varied, the reconstructions are very poor in the vanilla case, leading to poor tracking performance.}
\caption{\textcolor[rgb]{0,0,0}{Average number of successful valid sensory data transmissions per second versus duration of transmission phase $T_u$ under different number of UAVs. Distance between the BS and the tasks $=800$ m.}}
\caption{Broadband DOA estimation for $10$ sources {\color{\chang}(located at the angles $-66.9^\circ$, $-61.64^\circ$, $-42.84^\circ$, $-41.3^\circ$, $-5.74^\circ$, $-2.3^\circ$, $6.89^\circ$, $8.05^\circ$, $19.88^\circ$, $42.84^\circ$)} using an $m=15$ sensor ULA. The associated parameters are $k=10$, $q=100$, $d=5$, $c=3\times 10^8$, and $\sigma=1$. {\color{\change}The expected accuracy corresponding to three sets $\{\mathcal{P}_i\}_{i=1}^3$ are $\alpha_1=\frac{4}{5}$, $\alpha_2=\frac{2}{3}$, and $\alpha_3=0$. The top left, top right and bottom images correspond to the recovery using $\mathsf{P}_{1,2,\bm{D}\bm{\lambda}}^{\eta}$, when equal, heuristic and optimal weights $\bm{\lambda}$ (calculated using (\ref{eq.optimalweights_estimate})) are applied, respectively.}}
\caption{\label{tab:cascade:result} F1-measures and Accuracies of models on the test set of SARC$_{csd}$. The second three (4,5, and 6) models benefit from personality feature (their results are shown in \textcolor{blue}{blue}). Whereas the first three models (1,2, and 3), similar to our model; only rely on response or response and comment. Our models (AMR) achieves the F1-measure and accuracy of $68\%$ and $70\%$ respectively, the best results observed on SARC$_{csd}$ among similar methods which does not use personality features.}
\caption{System architecture with HotSpot components boxed in {\color{gray}gray} and our extensions for cloud index tracking in {\color{Maroon}red}.}
\caption{Precision / Recall plots showing the accuracy of BLL$_I$ (i.e., only individual hashtags), BLL$_S$ (i.e., only social hashtags), CF (i.e., Collaborative Filtering), \pred{} (i.e., individual and social hashtags) for $k$ = 1 - 10 recommended hashtags. \pred{} provides the best results with respect to predication accuracy but also fosters the reuse of hashtags and thus, confirmation bias and filter bubbles effects in our two Twitter datasets. \vspace{-3mm}}
\caption{\blue{(a-b): Spin integrated photoemission spectra of pristine and gold-covered Bi$_2$Se$_3$ using unpolarized He-I radiation ($h\nu\,{=}\,\SI{21.2}{eV}$). (c-e): Spin-resolved spectra using s-polarized laser light ($h\nu\,{=}\,\SI{6.0}{eV}$) collected for 0.3, 0.5 and 1.0~ML of gold. Dashed lines indicate the opening of a gap in the surface state. Further details are provided in the Supplementary Materials.}}
\caption{AGN duty cycle, i.e. percentage of cumulative time (upper panel) and cumulative energy (lower panel), above given Eddington ratio. The horizon dashed lines represent a fixed duty cycle of $50\%$. The lines in colors are the simulation results below given red shift. The points are observational constraints. The squares, circles, and upward- and downward- pointing triangles are from Ho (2009), Greene \& Ho (2007), Kauffmann\& Heckman (2009), and Heckman et al. (2004), respectively, which are all compiled from low-redshift observations. The star is a constraint compiled from high-redshift observations by Steidel et al. (2003).}
\caption{Performance evaluation for $15$ state-of-the-art algorithms on the VOT-2018 LTB35 dataset~\cite{VOTLT}. The best three results are marked in \textcolor{red}{\textbf{red}}, \textcolor{blue}{\textbf{blue}} and \textcolor{green}{\textbf{green}} bold fonts respectively. The trackers are ranked from top to bottom using the \textbf{F-score} measure.}
\caption{Ablation analysis of the proposed tracker on the VOT-2018 LTB35 dataset. The best results are marked in the \textcolor{red}{\textbf{red}} font. }
\caption{Detailed comparisons of different trackers on the OxUvA~\cite{OxUvA} dataset. The best three results are marked in \textcolor{red}{\textbf{red}}, \textcolor{blue}{\textbf{blue}} and \textcolor{green}{\textbf{green}} bold fonts respectively. The trackers are ranked from top to bottom using the \textbf{MaxGM} measure.}
\caption{(color online). (a) Evolution of the frequency of the in-plane normal modes of the 4-ion zigzag configuration as a function of the lattice vibrational frequency $\nu_{\textrm{latt}}$. As a reference the center of mass mode frequency of a single ion is shown by the gray dashed line. The two insets show the modes coordinates (in relative units) without lattice (0 MHz) and in presence of the deepest lattice (5 MHz). (b) Evolution of the coordinates of the modes 2 and 3 {\color{blue} (whose frequencies in absence of lattice are respectively the second and third lowest, and shown by the red and blue lines in (a))}, for low and high lattice frequencies (0 and 5 MHz) and around the avoided crossing (0.1, 0.15 and 0.2 MHz, indicated by the vertical dotted lines in (a)).}
\caption{Quantitative evaluation of state-of-the-art SR algorithms: average PSNR/SSIM for magnification factors $2\times$ and $4\times$. {\color{red} \textbf{Red}} indicates the best and {\color{blue} \underline{Blue}} indicates the second best performance. (`-' indicates that the method failed to reconstruct the whole images due to computation limitation.)}
\caption{An example radiology report with study background information organized into a \textbf{Background} Section, and radiology findings in a \textbf{Findings} Section. The human-written summary (or impression) and predicted summaries from different models are also shown. The extractive baseline does not summarize well, the baseline pointer-generator model generates \dred{spurious sequence}, while our model gives \dgreen{correct summary} by incorporating the background information.}
\caption{(a) Total stress as function of wall distance, normalized by dynamic pressure. (b,c,d) $\langle \f{u}_1'' \f{u}_1'' \rangle$, $\langle \f{u}_3'' \f{u}_3'' \rangle$, and $-\langle \f{u}_1'' \f{u}_3'' \rangle$ respectivelly, as function of wall distance. Linetypes ( {\color{red}---},{\color{blue}$--$},{\color{green}$-\cdot$},$\cdots$,{\color{cyan}--$+$}) correspond to different spacings.}
\caption{(continued on next page) Decay of Fourier modes of streamwise (left) and normal (right) dispersive components as function of wall distance $z/\delta$. (a) $S/\delta=0.3$ (b) $S/\delta=0.45$ ; (c) $S/\delta=0.88$; (d) $S/\delta=1.2$; (e) $S/\delta=1.8$. ({\color{red}---},{\color{blue}$--$},{\color{green}$-\cdot$},{\color{cyan}$\cdots$}): up to four Fourier modes with highest contribution to the dispersive shear component $\langle f{u}_1'' f{u}_3'' \rangle$; line types are ordered according to magnitude of the contributing mode. Lines are shifted with half a decade for visibility. (gray line in left plots): slope corresponding to $\exp(-2k_2z)$. (gray line in right plots): least-squares fit of $\exp(-2k_2z) (D_2 + D_3 z + D_4 z^2)$ to the data in the range $0.2<z<1.0$.}
\caption{\label{fig:calibration-landscape}A slice of the 3D calibration landscape for DRAG solution up to the first order in the small parameter to the qubit $\sigma_x$-gate leakage problem. Point A and B denote \cite{Motzoi2009}'s and \cite{Gambetta2011}'s first-order solutions, respectively. Point C is the optimum for this control function subspace (here $\alpha_x=-0.0069$), with infidelity of $10^{-6.63}$. A successful calibration process will typically start at a known DRAG solution, i.e. points A or B, and conclude in point C. The inset illustrates the associated pulse shapes: markers represent the unoptimized shapes ($u_x$: {\color{blue} $\bullet$}, $u_y$: {\color{red} $\blacksquare$}, $\delta$: {\color{green} $\blacklozenge$}) whereas solid lines depict the corresponding optimal solution (C).}
\caption{Intensity profile plotted as a histogram, with several characteristic positions: the mean value \textcolor{red}{VMEAN}, the median value \textcolor{blue}{VMEDIAN} and the maximum value\textcolor{green}{HMAX}}
\caption{{\color{blue}Average Dice scores and runtime results for affine alignment, ANTs, {\color{blue}NiftyReg} and \algname{} for the first experiment}. Standard deviations across structures and subjects are in parentheses. The average Dice score is computed over all structures and subjects. Timing is computed after preprocessing. Our networks yield comparable results to ANTs {\color{blue}and NiftyReg} in Dice score, while operating orders of magnitude faster during testing. We also show the number and percentage of voxels with a non-positive Jacobian determinant for each method, {{\color{bluedec}for our volumes with 5.2 million voxels within the brain}. All methods exhibit less than 1 percent such voxels.} \vspace{-0.25cm}}
\caption{{\color{blue}Boxplots of Dice scores for various anatomical structures for ANTs, {\color{blue}NiftyReg}, and VoxelMorph results for the first {\color{bluedec}(unsupervised)} experiment}. We average Dice scores of the left and right brain hemispheres into one score for this visualization. Structures are ordered by average ANTs Dice score. \vspace{-0.2cm}}
\caption{\color{blue}Example {\color{bluedec}deformation} fields $\bphi$ (columns 4-5) extracted by registering the moving image (column 1) to the fixed image (column 2) {\color{bluedec} in the unsupervised experiment (Section \ref{sec:main-expt}) }. The warped volume $\bmoving \circ \biphi$ is shown in column 3. {\color{bluedec}Displacement} in each spatial dimension is mapped to each of the RGB color channels in column 4. The {\color{bluedec}deformation} fields produced by VoxelMorph (MSE) are smooth within the brain, even when registering moving images that are significantly different from the fixed image. }
\caption{\color{blue} Effect of training set size on accuracy. Also shown are results of instance-specific optimization of deformations, after these are initialized with VoxelMorph outputs using the optimal global parameters resulting from the training phase.}
\caption{\color{bluedec}Results on test scans when using auxiliary data during training. Top: testing on the FreeSurfer segmentation of the general test set. Bottom: testing the same models on the manual segmentation of the Buckner40 test set. We test having varying number of observed labels (a-c), and having coarser segmentation maps (d). Error bars indicate standard deviations across subjects. The leftmost datapoint in each graph for all labels, corresponding to~$\gamma = 0$, indicates results of VoxelMorph without using auxiliary data (unsupervised). $\gamma = \infty$ is achieved by setting the image and smoothness terms to 0. We show Dice scores for results from ANTs with optimal parameters, which does not use segmentation maps, for comparison. \vspace{-0.5cm}}
\caption{\color{blue}Results for manual annotation experiment. We show affine, ANTs, NiftyReg, and \algname{}, where ``inst." indicates additional instance-specific optimization, as described in \textit{Section~\ref{sec:amortized}}. The average Dice score is computed over all structures and subjects, with standard deviations across structures and subjects in parentheses. }
\caption{\color{bluedec} Results for subject-to-subject alignment using affine, ANTs, and \algname{} variants, where ``x$2$'' refers to a model where we doubled the number of features to account for the increased inherent variability of the task, and ``inst." indicates additional instance-specific optimization. }
\caption{{\color{blue}Effect of $\gamma$ on warped images and deformation fields. We show the moving image, fixed image, and warped image (columns 1-3) with the structures that were observed at train time overlaid. The resulting deformation field is visualized in columns 4 and 5. While providing better Dice scores for observed structures, the deformation fields resulting from training with~$\gamma=\infty$ are far more irregular than those using~$\gamma=0.01$. Similarly, the warped image are visually less coherent for~$\gamma=\infty$. \vspace{-0.2cm}}}
\caption{\color{bluedec} Regularity of deformation fields when training with auxiliary segmentations obtained using FreeSurfer, MSE loss function and smoothness parameter of 0.02, measured using count and percentage of the number of voxels with non-positive Jacobian determinants.}
\caption{\color{blue}Example atlas-based VoxelMorph flow fields $\bphi$ (columns 4-5) extracted by registering the moving image (column 1) to the fixed image (column 2). The warped image $\bmoving \circ \biphi$ is shown in column 3. }
\caption{\color{blue}Auxiliary data experiment where the left and right hippocampus labels are observed at train time. We show the moving image, fixed image and warped image (columns 1-3) with the observed labels overlaid, and the resulting deformation fields (columns 4-5). We use the optimal $\gamma=0.01$ (left) and the extreme $\gamma=\infty$ (right). }
\caption{\color{blue}Auxiliary data experiment where a random half of the labels are observed at train time. We show the moving image, fixed image and warped image (columns 1-3) with the observed labels overlaid, and the resulting deformation fields (columns 4-5). We use the optimal $\gamma=0.01$ (left) and the extreme $\gamma=\infty$ (right). }
\caption{\color{blue}Auxiliary data experiment where all labels are observed at train time. We show the moving image, fixed image and warped image (columns 1-3) with the observed labels overlaid, and the resulting deformation fields (columns 4-5). We use the optimal $\gamma=0.01$ (left) and the extreme $\gamma=\infty$ (right). }
\caption{\color{blue}Auxiliary data experiment where coarse labels are observed at train time. We show the moving image, fixed image and warped image (columns 1-3) with the observed labels overlaid, and the resulting deformation fields (columns 4-5). We use the optimal $\gamma=0.01$ (left) and the extreme $\gamma=\infty$ (right). }
\caption[]{Visualization of distances in the descriptor space between all points on a selection of shapes from the FAUST test set and a single point \circlemarker on a reference shape (upper left). Cold colors correspond to small distances. Distances are saturated at the median.}
\caption[]{Comparison of descriptor performances on the FAUST test set. For four comparison methods \boxmarker{colorBlack} \boxmarker{colorRed} and our approach \boxmarker{colorBlue} we report the cumulative match characteristic (CMC), receiver operating characteristic (ROC) and the correspondence quality, which measures the distance between matched and ground truth points on the underlying mesh of the point cloud.}
\caption[]{Visualization of segmentation results on a selection of shapes from the FAUST test set. While the segmentation is visually convincing for most 3D point clouds, small inconsistencies can be observed. On the third shape in the top row points of the right lower arm \boxmarker{colorRLA}, right upper arm \boxmarker{colorRUA}, left lower arm \boxmarker{colorLLA} and left upper arm \boxmarker{colorLUA} are not always assigned to the correct side of the body. The same applies for points of the right upper leg \boxmarker{colorRUL} and left upper leg \boxmarker{colorLUL} on the fourth shape in the top row.}
\caption{Message sequence diagram for TLS~1.3 \texttt{Hello} messages with DHE key-exchange and \texttt{HelloRetryRequest}. Deprecated parameters that are included for backward compatibility are marked with \textcolor{gray}{gray} color.}
\caption{Contextual patterns. \textcolor{red}{Red} bounding boxes indicate the original labels provided by human annotators. \textcolor{blue}{Blue} bounding boxes indicate contexts mined by our algorithms. (a)~The \textit{contextual} bounding box is created by extending the small object region by $s=2$ pixels in horizontal and vertical directions. The information of the original labels~(red box) is preserved.; (b)~The \textit{augmented contextual} bounding boxes are created by extending or narrowing down the horizontal and vertical sides by $s=2$ pixels. Although there are more bounding boxes with diverse information, there are also many negative bounding boxes~(boxes inside the red frame) which do not preserve information of original label.}
\caption{\textbf{Cityscapes (Parsing $\rightleftharpoons$ Image).} The results for CycleGAN [\textcolor{green}{62}] are produced by the officially provided pretrained PyTorch models. GcGAN denotes GcGAN-\emph{rot}.}
\caption{\textbf{Google Mpas (Aerial photo $\rightleftharpoons$ Map).} For Map $\to$ Aerial photo, GcGAN produces competitive translations compared with CycleGAN [\textcolor{green}{62}].}
\caption{\textbf{SVHN $\to$ MNIST.} The qualitative results for both CycleGAN [\textcolor{green}{62}] and DistanceGAN [\textcolor{green}{5}] come from DistanceGAN [\textcolor{green}{5}]. The correct translations are about 24, 26, and 35 for CycleGAN, DistanceGAN, and GcGAN, respectively.}
\caption{\textbf{Horse $\to$ Zebra.} For this task, GcGAN generates slightly better translations for some images, but can not perform better than CycleGAN [\textcolor{green}{62}] generally.}
\caption{\textbf{Synthetic $\rightleftharpoons$ Real.} We train CycleGAN [\textcolor{green}{62}] using the released PyTorch codes. The results produced by GcGAN contain more details. Zoom in for better view.}
\caption{\small The illustration of multi-label image classification (MLIC) and weakly-supervised detection (WSD). % We show top-3 predictions, in which correct predictions are shown in \textcolor[rgb]{0.00,0.00,1.00}{blue} and incorrect predictions in \textcolor[rgb]{1.00,0.00,0.00}{red}. % The MLIC model might not predict well due to poor localization for semantic instances. % Although the detection results of WSD may not preserve object boundaries well, they tend to locate the semantic regions which are informative for classifying the target object, such that the predictions can still be improved.}
\caption{\small Example results on two datasets. The green bounding boxes in images are the top-10 proposals detected by T-WDet model, which is sorted by objectness confidences $s'$ in Eq. \ref{Eq:roi_feature}. The text on the right of images are the top-3 classification results of S-Cls model ``without'' and ``with'' knowledge distillation using our framework, where correct predictions are shown in \textcolor[rgb]{0.00,0.00,1.00}{blue} and incorrect predictions in \textcolor[rgb]{1.00,0.00,0.00}{red}.}
\caption{\small The improvements of S-Cls model over each class/concept on two datasets after knowledge distillation with our framework. ``\textcolor[rgb]{1.00,0.00,1.00}{*k}'' indicates the number (divided by 1000) of images including this class/concept. The classes/concepts in horizontal axis are sorted by the number ``\textcolor[rgb]{1.00,0.00,1.00}{*k}'' from large to small.}
\caption{$\mu_{\textrm{total}}$ results. The top performer for each experiment is highlighted in \textbf{bold}. Values within one standard deviation of the top performer are highlighted in {\color{blue}{blue}}. Note that when ExStream was best, no methods were within one standard deviation.}
\caption{$\mu_{\textrm{total}}$ results. The top performer for each experiment is highlighted in \textbf{bold}. Values within one standard deviation of the top performer are highlighted in {\color{blue}{blue}}. Note that when ExStream was best, no methods were within one standard deviation. }
\caption{Quantitative evaluation of our method compared to baselines on MSCOCO. \textcolor{blue}{\underline{Blue}} text indicates the performance after adjustments and \textbf{\textcolor{red}{red}} text indicates the best performance.}
\caption{Quantitative evaluation with our extension methods on MSCOCO. \textcolor{blue}{\underline{Blue}} text indicates the performance after adjustments and \textbf{\textcolor{red}{red}} text indicates the best performance.}
\caption{(a) Transition scenario as a function of particle concentration: Newtonian-like (\protect\markertwo), Particle-induced (\protect\markerone). (b) Friction factor at a fixed Reynolds number of $Re_s = 3500$ as a function of particle concentration. Experiments were done in the presence of the continuous perturbation.}
\caption{Two persistence diagrams overlaid, one from points sampled from a torus with two long-lived generators ({\color{blue}blue}) and the other from points sampled from an annulus with a single long-lived generator ({\color{red}red}). Some noisy (short-lived) generators have been truncated to simplify the example. The circled points and the edge between them highlights the pairing shown in \cref{fig:wass-graph-full} between $(5,1)$.}
\caption{Two persistence diagrams overlaid, one from a noisy torus with two long-lived generators ({\color{blue}blue}) and the other from a noisy annulus with a single long-lived generator ({\color{red}red}). Noisy (short-lived) generators have been truncated. The circled points and the edge between them highlights the pairing shown in \cref{fig:wass-example} between $(5,1)$.}
\caption{Illustration of differences among tree-based ensembles. The \textcolor{red}{red} rectangles show the union of the $5$ most anomalous subspaces across each of the $15$ most anomalous instances (\textcolor{blue}{blue}). These subspaces have the highest influence in propagating feedback across instances through gradient-based learning under our model. HST has fixed depth which needs to be high for accuracy (recommended $15$). IFOR has adaptive height and most anomalous subspaces are shallow. Higher depths are associated with smaller subspaces which are shared by fewer instances. As a result, feedback on any individual instance gets passed on to many other instances in IFOR, but to fewer instances in HST. RSF has similar behavior as HST. We set the depth for HST (and RSF) to $8$ (Figure~\ref{fig:hstrees_regions_8}) in our experiments in order to balance accuracy and feedback efficiency.}
\caption{Illustration of compact description and diversity using IFOR. Most anomalous $15$ instances (\textcolor{blue}{blue} checks) are selected as the query candidates. The \textcolor{red}{red} rectangles in {\bf (a)} form the union of the $\delta$ ($= 5$ works well in practice) most \emph{relevant} subspaces across each of the query candidates. {\bf (b)} and {\bf (c)} show the most ``compact'' set of subspaces which together cover all the query candidates. {\bf (b)} shows the most anomalous $5$ instances (\textcolor{green}{green} circles) selected by the greedy \texttt{Select-Top} strategy. {\bf (c)} shows the $5$ ``diverse'' instances (\textcolor{green}{green} circles) selected by \texttt{Select-Diverse}.}
\caption{\red{This table (if we need it) requires a caption. If we need the table we could add sketch of the patters and the loops???}}
\caption{A comparison of control flow APIs through an example data analysis program, which compares statistics of consecutive days from a year of page visit logs. \emph{PageAttributes} (marked with {\color{pageAttrColor} green}) is loop-invariant.}
\caption{Overview of the system architecture. Components that are the focus of this paper are {\color{blue} blue}. The concrete implementations of components are shown in [brackets], and are briefly discussed in \autoref{sec:impl}. }
\caption{Monopole of the redshift space galaxy correlation function of the main volume limited sample, with different corrections applied. The complete parent sample is shown in blue, targeted with no correction in yellow, assigning missing galaxies the redshift of the nearest targeted galaxy on the sky in green, transferring the weight of missing galaxies to the nearest targeted galaxy in red, angular upweighting in purple, and PIP weighting in brown. \red{The ratio to the complete parent sample, for different correction methods, is split between the two lower panels for clarity.} Shaded regions are errors estimated from 100 jackknife samples. Horizontal black dotted lines indicate $\pm 1 \%$. For $s \gtrsim 20 \hMpc$, the scatter is almost the same for all methods.}
\caption{The exchange/correlation energy $E_{x/c}$, vs. Wigner-Seitz radius $r_s$, shown on a log-log plot, for \textcolor{jj}{the $N_e$-electron gas ($N_e$-EG)} with $N_{e}=2$, 3, \& 4 electrons. Shown also are literature results for the\textcolor{BP}{UEG system, obtained from QMC [Refs.~\cite{RN819} and \cite{RN667}] and related DFT [Ref.~\cite{RN658}] simulations.} %\textcolor{jj}{, which are in the many-electron limit}. For each \textcolor{jj}{$N_e$-EG} system, the colored region corresponds to the range of values whose upper and lower bounds are determined by enforcing exact spin symmetry on the $N_{e}$-electron wave-function states used to compute $E_{x/c}$: 2-\textcolor{jj}{EG} (blue region); 3-\textcolor{jj}{EG} (gray region); and 4-\textcolor{jj}{EG} (red region). The black lines correspond to the fully unrestricted exchange/correlation energies. %Numerical values for the ground-state energy and correlation energy vs. $r_s$ is included in the SI. %\textcolor{BP}{!! WHAT IS VWN? SHOULD WE DEFINE ACRONYM?} }
\caption{Fermi energy $\mu$ \textcolor{BP}{for few-EG systems}, in both the high-density (a) and low-density (b) regimes. Note the change of scale from log-log in (a) to log-linear in (b). The dashed curves in (a) correspond to the non-interacting limit of $\mu \propto r_s^{-2}$ for a system dominated by the kinetic energy contribution. The color schemes are the same in both (a) and (b). %\textcolor{BP}{REPLACE LABELS IN (A) TO BE 4-EG, ETC.} }
\caption{Electron/electron pair-correlation functions $g(r)$, for \textcolor{jj}{few-EG systems} with 2 (a), 3 (b), and 4 (c) electrons, for various $r_s$ values ranging from $r_s = 1$ to 1024au. Frame (d) compares $g(r)$ for all three $N_e$ cases, for a Wigner-Seitz radius $r_s=256$au lying squarely in the BCC Wigner crystal regime. }
\caption{ Electron occupation numbers (by permutation group irrep) vs. $r_s$, for 2-,3-, and 4-\textcolor{jj}{EG systems}. Irreps refer to spatial symmetry. Each graph is offset for clarity, with the 0-baseline for each indicated as a dashed line. For the 2- and 3-\textcolor{jj}{EG cases} %cases , only two spatial symmetries are permitted by the Exclusion Principle: $A_1$ and $A_2$, and $A_2$ and $E$. %JJ, respectively, of which only $A_2$ is represented here. The 4-\textcolor{jj}{EG case} %case has 3 allowed irreps, $A_2$, $E$, and $T_1$, all of which are represented here. }
\caption{\textcolor{BP}{The three parameters of the semi-empirical liquid drop model (see Eq.~\ref{ldm}), as a function of Wigner-Seitz radius, $r_s$: volume (magenta); area (green); curvature (orange).} The three parameter curves have been fit to power laws for the higher density regime $(r_s<10)$, as indicated in the figure. The relative smoothness of these plots suggests \textcolor{BP}{the validity of} the liquid drop model \textcolor{BP}{throughout the $r_s$ range.} %\textcolor{jj}{The unmarked lines are the standard QMC results\cite{RN819}} %in this context; however, the volume term is far from the UEG limit. %\textcolor{BP}{!! How far? There is no discussion of this in paper} }
\caption{comparison of AUC scores between DGCs and the corresponding baselines on ChestX-ray14 dataset. The better results based on the same encoder are \textbf{bolded}. The best results among all the models are colored \textcolor{red}{red}.}
\caption{\textbf{Challenge results.} The top 9 submissions in each region. For submissions with a marginal PI difference (up to 0.01), the one with the lower RMSE is ranked higher. Submission with marginal differences in both the PI and RMSE are ranked together (marked by $*$). We perform a human-opinion-study on the {\color{blue}top submissions} colored in blue (see Section \ref{sec:humanStudy}). See the cited papers describing the submissions. Team members and affiliations can be found in Appendix \ref{ap:teamMembers}. A full table of the test phase results appears in Appendix \ref{app:fullResults}.}
\caption{(color online) Ground-state energies for \elem{He}{4} obtained in the NCSM with the HO, HF, and NAT basis sets (panels from left to right) as function of the oscillator frequency $\hbar\Omega$ for $N_{\max}=4$ (\symbolcircle[FGBlue]), $6$ (\symboldiamond[FGRed]), $8$ (\symboltriangle[FGGreen]), $10$ (\symbolbox[FGViolet]), and $12$ (\symbolcross[FGLightBlue]). All calculations employ the chiral NN+3N interaction ($\Lambda_{\text{3N}}=400\,\text{MeV}/c$) after an SRG evolution with $\alpha = 0.08\,\text{fm}^4$.}
\caption{(color online) Ground-state energies for \elem{O}{16} obtained in the NCSM with the HO, HF, and NAT basis sets with $N_{\max}=2$ (\symbolcircle[FGBlue]), $4$ (\symboldiamond[FGRed]), $6$ (\symboltriangle[FGGreen]), $8$ (\symbolbox[FGViolet]), and $10$ (\symbolcross[FGLightBlue]). The open symbols indicated results obtained with the NO2B approximation. All other paramters as in Fig.~\ref{fig:He4egs}.}
\caption{(color online) Point-proton radii for the ground state of \elem{He}{4} obtained in the NCSM with the HO, HF, and NAT basis sets (panels from left to right) as function of the oscillator frequency $\hbar\Omega$ for $N_{\max}=4$ (\symbolcircle[FGBlue]), $6$ (\symboldiamond[FGRed]), $8$ (\symboltriangle[FGGreen]), $10$ (\symbolbox[FGViolet]), and $12$ (\symbolcross[FGLightBlue]). All calculations employ the chiral NN+3N interaction ($\Lambda_{\text{3N}}=400\,\text{MeV}/c$) after an SRG evolution with $\alpha = 0.08\,\text{fm}^4$.}
\caption{(color online) Point-proton radii for the ground state of \elem{O}{16} obtained in the NCSM with the HO, HF, and NAT basis sets with $N_{\max}=2$ (\symbolcircle[FGBlue]), $4$ (\symboldiamond[FGRed]), $6$ (\symboltriangle[FGGreen]), $8$ (\symbolbox[FGViolet]), and $10$ (\symbolcross[FGLightBlue]). All other paramters as in Fig.~\ref{fig:He4rad}.}
\caption{(color online) Ground-state energies of all oxygen isotopes from \elem{O}{14} to \elem{O}{26} using the chiral NN+3N interaction ($\Lambda_{\text{3N}}=400\,\text{MeV}/c$) after an SRG evolution with $\alpha = 0.08\,\text{fm}^4$. Shown are NCSM calculations using the NAT basis ($\hbar\Omega=20\,\text{MeV}$) with the explicit 3N interaction (\symbolbox[FGBlue]) and with the NO2B approximation (\symboldiamond[FGRed]). For comparison we also plot the HO basis calculations of Ref.~\cite{HeBi13} with optimized oscillator frequency (\symbolcircle[FGGreen]). The inset shows the $N_{\max}$ dependence and extrapolation for selected isotopes with the explicit 3N interaction (see text).}
\caption{Evolution of training loss and training accuracy per batch for all four centrality metrics throughout the training process. The loss is plotted in {\color{Bittersweet}orange} and {\color{blue}blue} and the accuracy in {\color{red}red} and {\color{ForestGreen}green} for training with and without multitasking, respectively.}
\caption{Evolution of the 1D projection of vertex embeddings plotted against the corresponding eigenvector centralities through time for a graph sampled from the Watts-Strogatz small world distribution. Vertices are coloured according to their eigenvector centrality rank (the most central elements are {\color{blue}blue} and the least central ones are {\color{red}red})}
\caption{Evolution of the 2D projection of vertex embeddings through time for a graph sampled from the power law cluster distribution. Vertices are coloured according to their eigenvector centrality rank (the most central elements are {\color{blue}blue} and the least central ones are {\color{red}red})}
\caption{Universal decrease of Fermi-velocity anisotropy for long-range interacting Dirac fermions. Analytical perturbation theory and Hartree-Fock results for Dirac fermions (red dashed line) show a square-root decrease of the Fermi velocity anisotropy. Lattice perturbation theory on the honeycomb lattice (green diamonds){\color{highlight},} the non-perturbative projective Quantum Monte Carlo {\color{highlight} on honeycomb lattice} (purple triangles) {\color{highlight} and $\pi$-flux model (red circles)} are {\color{highlight} all} consistent with this square-root behaviour. Also shown are recent experiments (black squares) and density matrix renormalization group calculations (blue circles) for the $\nu = 1/2$ quantum Hall state. We find that both the chirality of the Dirac bands and a long-range interaction potential are necessary for this universal decrease in anisotropy. \label{fig:Fig1}}
\caption{Non-perturbative quantum Monte Carlo simulations on a {\color{highlight} $\pi$-flux model and the honeycomb lattice both} show universal square-root anisotropy renormalization with long-range Coulomb interactions, but not for Hubbard interactions. (a) QMC data {\color{highlight} on honeycomb lattice} with onsite Hubbard interaction (left panel) and long-range Coulomb interaction (right panel). The energy renormalization $E-E_0$ of the quasiparticle is obtained for $K-\Gamma$ direction (squares) and $K-K'$ direction (triangles). Solid lines indicate what one would expect for a square-root renormalization, while dashed lines show no anisotropy renormalization. (b) The interacting theory anisotropy as a function of momentum away from the Dirac point. The {\color{highlight} $\pi$-flux model has an anisotropy that is almost constant for small momentum down to the Dirac point, while the} honeycomb lattice is isotropic as momentum vanishes but becomes more anisotropic at larger momentum. Both the non-perturbative QMC and the lattice perturbation theory show a clear square-root anisotropy renormalization for long-range Coulomb interactions. \label{fig:Fig4}}
\caption{\red{Comparisons of the atmospheric parameters estimated from ECMWF and measured by stereo-SCIDAR at ESO Paranal. (a) is the seeing, (b) the free atmosphere seeing, (c) the ground layer seeing, (d) the coherence time and (e) is the for the isoplanatic angle.}}
\caption{\red{Comparisons of the distribution of atmospheric parameters estimated from ECMWF and measured by stereo-SCIDAR at ESO Paranal. (a) is for the total integrated seeing. It can be seen that in this case the ECMWF model and the stereo-SCIDAR measurements share a very similar distribution. (b) shows the comparison for the free atmosphere only ($h$>1~km above observatory level or 3.6~km above observatory level). In this case the the ECMWF model does not show the variability of the free atmosphere seeing that is measured by the stereo-SCIDAR, particularly for the more turbulent conditions. (c) shows the distribution of seeing values integrated up to 1~km. The distribution of the coherence time and the isoplanatic angle are shown in (d) and (e) respectively. The coherence time shows good agreement, however the isoplanatic angle from the ECMWF model does not show the same variability of values as the stereo-SCIDAR measurements.}}
\caption[Data-set characteristics, sagittal foot orientation, by movement type and stance limb]{\color{caption_main}\textbf{Data-set characteristics, sagittal foot orientation, by movement type and stance limb.}}
\caption[CaffeNet CNN model training (double-cascade)]{\color{caption_main}\textbf{CaffeNet CNN model training (double-cascade).}~{\color{caption_sub}Input marker trajectories flattened to color images, output KJM 6-way deinterlaced PCA reduced.}}
\caption[Training-set eight marker trajectories, sidestep~left movement type (off right stance limb)]{\color{caption_main}\textbf{Training-set eight marker trajectories, sidestep~left movement type (off right stance limb).}~{\color{caption_sub}Combined 1,222 predictor samples (80~\% of 1,527), viewed as a conventional 3D volume space and with one trial highlighted.}}
\caption[KJM component vector and mean correspondence, CNN single fine-tune]{\color{caption_main}\textbf{KJM component vector and mean correspondence $r(rRMSE~\%)$, CNN single fine-tune.}~{\color{caption_sub}33~\%~stance, by movement type and stance limb. CaffeNet CNN, output 6-way deinterlaced PCA reduced, single 80:20 fold.}}
\caption[KJM component vector and mean correspondence, CNN double-cascade]{\color{caption_main}\textbf{KJM component vector and mean correspondence $r(rRMSE~\%)$, CNN double-cascade.}~{\color{caption_sub}33~\%~stance, by movement type and stance limb. CaffeNet CNN, output 6-way deinterlaced PCA reduced, single 80:20 fold.}}
\caption[Ground truth versus predicted response]{\color{caption_main}\textbf{Ground truth versus predicted response.}~{\color{caption_sub}Top, training-set modeled output sidestep~left (right stance foot) internal moments $RKJM_x$ extension/flexion, $RKJM_y$ abduction/adduction, and $RKJM_z$ internal/external rotation, 1,222 samples versus stance phase. Middle and lower, test-set modeled internal KJM (blue, ticks), and predicted response (red), using CaffeNet double-cascade, 6-way deinterlaced, correlations over initial 33~\% stance phase. Middle, min/max range and mean predicted response; lower, individual sample with the strongest $r(RKJM_{mean})$.}}
\caption{Surface pressure coefficient $C_p$ as a function of the angle $\theta$ ($\theta=0^\circ$ corresponds to the stagnation point, $\theta=180^\circ$ to the base point) for the case of a stationary cylinder in uniform flow at $Re_D=100$. $\bullet$, present results; \solid, data from Park et al.~\cite{park:98}.}
\caption{ Drafting-kissing-tumbling of two sedimenting discs with density $\rho_p/\rho_f=1.5$ which are initially aligned vertically. Results obtained with the present method, $\Delta t=0.0001$, $\Delta x=1/256$: {\solid trailing}, {\dotted leading}. Results provided by T.-W.\Pan:{\chndot trailing}, {\dashed leading}. Vertical position (left), vertical velocity (right). The interval during which the repulsion force takes finite values is indicated near the abscissa. }
\caption{ Drafting-kissing-tumbling of two sedimenting discs with density $\rho_p/\rho_f=1.5$ which are initially aligned vertically. Results obtained with the present method, $\Delta t=0.0001$, $\Delta x=1/256$: {\solid trailing}, {\dotted leading}. Results provided by T.-W.\Pan:{\chndot trailing}, {\dashed leading}. Horizontal position (left), horizontal velocity (right). }
\caption{ Rotational velocity vs.\time during the interaction of two sedimenting discs with density$\rho_p/\rho_f=1.5$ which are initially aligned vertically. Results obtained with the present method, $\Delta t=0.0001$, $\Delta x=1/256$: {\solid trailing}, {\dotted leading}; results provided by T.-W.\Pan:{\chndot trailing}, {\dashed leading}. }
\caption{Wake interaction of two sedimenting particles with density $\rho_p^{(1)}/\rho_f=1.5$, $\rho_p^{(2)}/\rho_f=1.25$ and initial vertical and lateral offset. The gravity acts in the positive $x$-direction. Results obtained with $\Delta t=0.001$, $h=1/200$. The graph shows the particle trajectories with the line stlyes corresponding to: \solid~{heavy particle}, \dashed~light particle.}
\caption{Sedimentation of a single sphere; case 1 of reference~\cite{mordant:00}. Vertical velocity: \solid present, \dashed experimental data.}
\caption{Sedimentation of a single sphere; case 2 of reference~\cite{mordant:00}. Vertical velocity: \solid present, \dashed experimental data.}
\caption{Sedimentation of a single sphere; case 4 of reference~\cite{mordant:00}. Vertical velocity: \solid present, \dashed experimental data.}
\caption{Sedimentation of a single sphere. Horizontal velocities $u$ (left) and $v$ (right). \solid~case~1, \dashed~case~2, \dotted~case~4.}
\caption{Sedimentation of an array of $N_p$ identical spheres with density ratio $\rho_p/\rho_f=2.56$ and terminal Reynolds number approximately $400$. Mean vertical particle velocity vs.\time for:\dotted~case~1~($N_p$=$1$), \dashed~case~2~($N_p$=$63$), \solid~case~3~($N_p$=$1000$).}
\caption{Sedimentation of an array of $n$ identical spheres with density ratio $\rho_p/\rho_f=2.56$ and terminal Reynolds number approximately $400$. Mean distance to the nearest particle vs.\time for:\dashed~case~2~($N_p$=$63$), \solid~case~3~($N_p$=$1000$).}
\caption{A sample chat from our dataset which uses background resources. The chosen spans used in the conversation are shown in \textcolor{blue}{\uline{blue}}. The letters in the brackets denote the type of resource that was chosen - P, C ,R, F and N indicate \textbf{P}lot, \textbf{C}omments, \textbf{R}eview, \textbf{F}act Table and \textbf{N}one respectively.}
\caption{\label{fig} The (real part of the) ghost propagator $G_{\text{r}}(p)$ in minimal LCG (\protect\marksymbol{square*}{red}) and in Landau gauge (\protect\marksymbol{*}{black}), as a function of the lattice momentum $p$, with $p_{\mu}(k) = 2 \sin{(\pi k_{\mu}/N)}$ and $k_{\mu} = 1,2,\ldots,N/2$. Note the logarithmic scale on the $y$ axis. Both $G_{\text{r}}(p)$ and $p$ are in physical units. {\bf Top}: SU(2) case with $V = 24^4$, $\beta = 2.4469$ and $\xi = 0.1$, corresponding to the continuum value $\hat{\xi} = 0.24469$, for 60 thermalized configurations. {\bf Bottom}: SU(3) case with $V = 24^4$, $\beta = 6.0$ and $\xi = \hat{\xi} = 0.1$ for 79 thermalized configurations. }
\caption{The Gaussian distributions in {\color{blue} blue} and {\color{Gray} gray} are our estimations. The Dirac delta function in {\color{Orange} orange} is the distribution of the ground-truth bounding box. When the location $x_e$ is estimated inaccurately, we expect the network to be able to predict larger variance $\sigma^2$ so that $L_{reg}$ will be lower (blue)}
\caption{\ournms}{$\mathcal{B}$ is $N\times4$ matrix of initial detection boxes. $\mathcal{S}$ contains corresponding detection scores. $\mathcal{C}$ is $N\times4$ matrix of corresponding variances. $\mathcal{D}$ is the final set of detections. $\sigma_t$ is a tunable parameter of \ournms. The lines in {\color{blue}blue} and in {\color{ForestGreen}green} are soft-NMS and \ournms\respectively.}
\caption{Results for TLS parameter mapping/mirroring and vulnerabilities to known attacks. For TLS version mapping, we display the TLS versions seen by the client when the web server uses TLS 1.2, 1.1, 1.0 and SSL 3.0 (`--' means unsupported. `\textdagger' means supported but terminate with a handshake failure; see Section~\ref{tlspararesults}). Under ``Key Length Mapping'': `*' means the appliance mirrors RSA-512 and RSA-1024 key sizes, but use a static key size RSA-2048 for any higher key sizes (see Section~\ref{tlsversionmapping}). Under ``Problematic Ciphers'': Weak means deprecated; Insecure means broken; blank means good ciphers. Under ``BEAST'': \rx \textcolor{black}{} means vulnerable; \rx\red{*} \color{black}{}means potentially vulnerable (unknown if CBC is patched with $1/(n-1)$ split); blank means patched. All the appliances are patched against FREAK, Logjam, CRIME, and Insecure Renegotiation.}
\caption{\label{fig:comp}\textcolor{gray}({Color online) Material distribution in the system, in the case of lens- (a) and disk-shaped (b) QD. } }
\caption{\label{fig:mag_h}\textcolor{gray}({Color online) Magnetic-field dependence of the lowest hole energy levels for the lens shaped [001]-oriented QD. The inset contains enlarged part of the plot with anticrossing between $s$- and $p$-type states. Energy $E=0$ refers to the unstrained GaAs valence-band edge. } }
\caption{\label{fig:cqd}\textcolor{gray}({Color online) Magnetic-field dependence of the lowest hole energy levels for the disk shaped [001]-oriented QD. The inset contains the avoided crossing width between $s$- and $p$-shell energy levels as a function of external axial electric field $F$. } }
\caption{\label{fig:qd111}\textcolor{gray}({Color online) Magnetic field dependence of the lowest hole energy levels for the lens shaped [111]-oriented QD. } }
\caption{\label{fig:ph}\textcolor{gray}({Color online) (a) Phonon-assisted spin relaxation rate in the lowest-energy Zeeman doublet as a function of magnetic field for [$001$]- and [$111$]-oriented QDs; (b) ratio of the relaxation rate from spin-admixture mechanisms to the overall relaxation rate. } }
\caption{\label{fig:harm}\textcolor{gray}({Color online) Form-factor $F(q)$ expansion in spherical harmonics for (a) [$001$]-oriented lens- and (b) disk-shaped QD, and (c) [$111$]-oriented lens-shaped QD.} }
\caption{\label{fig:tun-harm}\textcolor{gray}({Color online) Phonon-assisted spin relaxation rate due to coupling via piezoelectric potential; red points denotes results obtained taking numerically exact $F(q)$, for the black solid line $F(q) \approx a_{0,0} Y^0_0 $, and for the blue dashed line $F(q) \approx a_{1,-1} Y^{-1}_1 + a_{1,0} Y^{0}_1 + a_{1,1} Y^{1}_1$.} }
\caption{Left column uniform sampling. Right column nonuniform sampling. (\subref{fig:results_PSD:uniform:Euclidean_sampling})-(\subref{fig:results_PSD:nonuniform:Euclidean_sampling}) Intrinsically stationary signal in Euclidean space with $\gamma(h) = 1 - \exp \big(-\frac{h}{0.2}\big)$ and sampled on the graph vertex locations. (\subref{fig:results_PSD:uniform:PSD})-(\subref{fig:results_PSD:nonuniform:PSD}) empirical graph PSD as a function of normalized graph frequency, $\lambda$. Note that for both sampling schemes high correlations for short distances ($\gamma(0) = 0$) lead to low energy in the higher spectrum since higher graph frequencies correspond to large variations on edges with large weights. } \label{fig:results_PSD} \end{figure} %In this section, we describe our approach extending the variogram to the graph setting leading to our definition of graph intrinsic stationarity. \subsection{Relating the Variogram to the Graph Laplacian} In the context of sensor networks, we have direct access to pairwise distances between nodes, which allows us to introduce the concept of the variogram. For our problem formulation we use the \textit{graph Laplacian quadratic form} of a graph signal $\mathbf{x}$: \begin{align} \mathbf{x^{T}Lx} = \frac{1}{2}\sum_{i,j = 1}^{N} w_{ij}\big[ x_i - x_j \big]^{2} = \sum_{\mathclap{(i,j) \in E}} w_{ij}\big[ x_i - x_j \big]^{2}, \label{eq:quadform_original} \end{align} where $\mathbf{x^{T}Lx}$ %as defined in \eqref{eq:quadform_original} measures squared differences of random field values across all edges $(i,j) \in E$. For constant equal weights $w_{ij} = 1$, this quadratic form measures the variation of the graph signal on a global-scale. Comparing \eqref{eq:local_variogram_empirical} and \eqref{eq:quadform_original}, we see that the empirical isotropic local variogram, $2 \hat{\gamma}(h;\mathbf{s_k})$, is a valid graph Laplacian quadratic form. More precisely, the normalized weights $\smash{\frac{\mathrm{W}(\mathbf{s_i},\mathbf{s_j};h;\mathbf{s_k})}{\mathrm{W}(h;\mathbf{s_k}) }}$ play the same role as that of $w_{ij}$ in \eqref{eq:quadform_original}. Computing the generalized empirical isotropic local variogram is therefore equivalent to computing the quadratic form \eqref{eq:quadform_original} for a graph Laplacian matrix that best describes $\mathrm{N}_{r}(h;\mathbf{s_k})$. %Next we %We now have a simple alternative for computing empirical variograms on an irregular graph domain. The following section %describe how to define the appropriate $\mathbf{L}$ for computing the empirical isotropic local variogram for $N_{r}(h;\mathbf{s_k})$. \subsection{Defining \texorpdfstring{$\mathrm{N}_{r}(h;\mathbf{s_k})$}{Nr(h;sk)} on a Graph} %The graph Laplacian quadratic form method for computing $2 \hat{\gamma}(h;\mathbf{s_k})$ requires using a new Laplacian matrix with a set of weights that satisfy \eqref{eq:window_separable} where $w_h(\cdot)$ is accounted for. Though they are similar, the key difference between \eqref{eq:local_variogram_empirical} and \eqref{eq:quadform_original} is that edge weights $w_{ij}$ in \eqref{eq:quadform_original} correspond to arbitrary distances. This motivates the need to bin distances in order to make \eqref{eq:quadform_original} closely match the definition of the empirical isotropic variogram in \eqref{eq:local_variogram_empirical}. This amounts to creating a different Laplacian for each $(h,\mathbf{s_k})$. To make \eqref{eq:quadform_original} match more closely with \eqref{eq:local_variogram_empirical}, we first define a new adjacency matrix $\mathbf{{A}_{\boldsymbol{\Delta_h}}}$, that assigns zero weights to edges with corresponding distance length outside prespecified bin width tolerance range ${\Delta_h} \coloneqq (h - \frac{\delta}{2},h +\frac{\delta}{2} )$. This reduces to applying an elementwise binary operator on the original adjacency matrix. %Following the definition of $\mathrm{N_r} (h,\mathbf{s_k})$used in \eqref{eq:binary_local_variogram_empirical}, we design a new adjacency matrix For any pair of nodes $(i,j)$: \begin{align} [ \mathbf{A_{\boldsymbol{\Delta_h}}}]_{i,j} & = \left\{ \begin{array}{cc} 1, & \hspace{1mm} d_{ij} \in {\Delta_h} \\ 0, & \hspace{1mm} d_{ij} \notin {\Delta_h} \end{array} \right. \label{eq:adjacency_previous} \end{align} The corresponding graph Laplacian $\mathbf{{L}_{\boldsymbol{\Delta_h}}}$ is calculated from $\mathbf{{A}_{\boldsymbol{\Delta_h}}}$. Suppose out of all pairs of nodes, the maximum distance is $d_{max}$. ${\Delta_h} $ is therefore contained within $(0,d_{max})$. We propose here to break the interval $(0,d_{max})$ into $H$ mutually disjoint intervals $\{ {\Delta_h} \}^{H}_{h=1}$ to obtain corresponding Laplacians $\{ \mathbf{{L}_{\boldsymbol{\Delta_h}}} \}^{H}_{h=1}$. We will study alternative binning methods in a future communication. By this stage, using the isotropic assumption we have built the necessary tools to compute the global variogram shown in \eqref{eq:empirical_variogram}. Alternatively, we can compute the generalized empirical local variogram shown in \eqref{eq:local_variogram_empirical}. To do so, we use non-binary windows centered on vertex $k$ decaying with distance to $k$. We propose using a graph signal $\mathbf{g_{k}} \in \mathbb{R}^{N}$ localized at node $k$, to play the role of a local window in the vertex domain. Using $\mathbf{g_{k}}$ to modify our previous adjacency matrix in \eqref{eq:adjacency_previous}, we write $\mathbf{{A}}_{\boldsymbol{(\Delta_h,k)}}$: \begin{align} \mathbf{{A}}_{\boldsymbol{(\Delta_h,k)}} = \mathbf{G_{k}A_{\boldsymbol{\Delta_h}}}\mathbf{G_{k}} &\quad \text{with} \quad& \mathbf{G_k} = \mathrm{diag}\big( \mathbf{g_k} \big) \end{align} The normalization constant that would be analogous to $W(h;\mathbf{s_k})$ in \eqref{eq:normalization_term} is accounted for by computing the quadratic form of the degree matrix $\mathbf{D}_{\boldsymbol{(\Delta_h,k)}}$: \begin{align} \mathbf{{W}}_{\boldsymbol{(\Delta_h,k)}} = \mathbf{1}^{T} \mathbf{{D}}_{\boldsymbol{(\Delta_h,k)}}\mathbf{1} \label{eq:degree_normalization} \end{align} Here $\mathbf{D}_{\boldsymbol{(\Delta_h,k)}}$ accounts for degree variations as a function of both $\Delta_h$ and $k$. \subsection{Graph Variogram} We can now proceed with writing our final expression for calculating the GSP-based empirical isotropic local variogram. We call this quadratic form the graph local variogram $2\gamma_{G}$, where $2\gamma_{G}({\Delta_h},k)$ defines an empirical local variogram relative to center node $k \in V$ with respect to distance bin $\Delta_h$: \begin{align} 2\gamma_{G}({\Delta_h},k)= 2 \frac{\mathbf{x}^{T} \mathbf{{L}}_{\boldsymbol{(\Delta_h,k)}}\mathbf{x}}{\mathbf{1}^{T} \mathbf{{D}}_{\boldsymbol{(\Delta_h ,k)}}\mathbf{1}} \label{eq:local_graph_variogram} \end{align} where $ \mathbf{{L}}_{\boldsymbol{(\Delta_h , k)}} = \mathbf{{D}}_{\boldsymbol{(\Delta_h , k)}}- \mathbf{{A}}_{\boldsymbol{(\Delta_h , k)}}$. $2\gamma_{G}({\Delta_h},k)$ is undefined whenever the neighborhood described by ($\Delta_h ,\mathbf{s_k}$) is empty. The 2 in the r.h.s. of \eqref{eq:local_graph_variogram} is included to account for the quadratic Laplacian form in \eqref{eq:quadform_original} using each edge only once, whereas the original variogram in \eqref{eq:local_variogram_empirical} counts each edge twice. We further define the graph global variogram as the average over vertices of the graph local variogram: \begin{align} 2\gamma_{G}(\Delta_h) = \frac{1}{N}\sum_{k=1}^{N} 2\gamma_{G}(\Delta_h , k ) \end{align} $2\gamma_{G}(\Delta_h)$ measures spatial variations on a global scale for given ${\Delta_h}$. Finally, we propose the following definition of intrinsic global graph stationarity: \textbf{Definition:} (Intrinsic Global Graph Stationarity) \textit{A signal $\mathbf{x}$ defined over G(V,E) is intrinsically stationary if and only if } \begin{align} \mathbb{E}\big[\gamma_{G}(\Delta_h ,k)\big] &= \mathbb{E}\big[ \gamma_{G}(\Delta_h)\big], \forall k \in V \end{align} Intuitively, this definition states that no matter the center vertex $k$, the local graph variogram is the same.% Therefore, this definition implements a shift invariance in the vertex domain that is closer to the classical definition of stationarity, as opposed to graph stationarity, which implements a shift invariance in the spectral domain \cite{BG1}. \section{Results} \label{sec:experiments} To validate the definition of the global variogram, we now perform experiments showing closeness to the theoretical true variogram. More precisely, we compute realizations of an intrinsic stationary field with an isotropic true variogram 2$\gamma(h)$. Simultaneously, we generate an independent sampling of the field for each realization of the isotropic model with the goal of showing robustness to both sources of noise. Both uniform and non-uniform samplings (see \autoref{fig:results_PSD:uniform:Euclidean_sampling} and \autoref{fig:results_PSD:nonuniform:Euclidean_sampling}) have been used within a square grid defined by $\{\forall (x,y) \in \mathbb{R}^{2} : 0\leq x \leq1, 0\leq y \leq1 \}$ as shown in the following set of figures, which were generated using GraSP \cite{BG2}. We build the original graphs using two schemes: fully connected and sparse using $K$-nearest neighbors ($K$ =100). In both cases, the edge weights are chosen using a Gaussian kernel of the Euclidean distance with parameter $\sigma = 0.05$. \begin{figure}[tb] \centering \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure3.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% %ylabel={$PSD_{\infty}(\lambda)$}, %legend entries={$\mu_{ PSD_{\infty}}\pm \sigma_{ PSD_{\infty}}$,$\mu_{ PSD_{\infty} }$}, %legend style={at={(0.5,-0.6)},anchor=north}, %legend style={font=\tiny}, legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\gamma_{\infty}}$,$\mu_{\gamma_{\infty}}$,$\gamma$}, legend pos=south east,% reverse legend ] \pgfplotstableread{U_NOKNN_data_singlegraph.dat}{\UNOKNNSINGLE} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\UNOKNNSINGLE}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\UNOKNNSINGLE}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\UNOKNNSINGLE}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:uniform:noknn_singlegraph} \end{subfigure} %% \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure4.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% %ylabel={$PSD_{\infty}(\lambda)$}, %legend entries={$\mu_{ PSD_{\infty}}\pm \sigma_{ PSD_{\infty}}$,$\mu_{ PSD_{\infty} }$}, %legend style={at={(0.5,-0.6)},anchor=north}, %legend style={font=\tiny}, legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\gamma_{\infty}}$,$\mu_{\gamma_{\infty}}$,$\gamma$}, legend pos=south east,% reverse legend ] \pgfplotstableread{NU_NOKNN_data_singlegraph.dat}{\NUNOKNNSINGLE} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\NUNOKNNSINGLE}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\NUNOKNNSINGLE}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\NUNOKNNSINGLE}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:nonuniform:noknn_singlegraph} \end{subfigure} \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure5.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% %ylabel={$PSD_{\infty}(\lambda)$}, %legend entries={$\mu_{ PSD_{\infty}}\pm \sigma_{ PSD_{\infty}}$,$\mu_{ PSD_{\infty} }$}, %legend style={at={(0.5,-0.6)},anchor=north}, %legend style={font=\tiny}, legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\gamma_{100}}$,$\mu_{\gamma_{100}}$,$\gamma$}, legend pos=south east,% reverse legend ] \pgfplotstableread{U_KNN_data_singlegraph.dat}{\UKNNSINGLE} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\UKNNSINGLE}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\UKNNSINGLE}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\UKNNSINGLE}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:uniform:knn_singlegraph} \end{subfigure} \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure6.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% %ylabel={$PSD_{\infty}(\lambda)$}, %legend entries={$\mu_{ PSD_{\infty}}\pm \sigma_{ PSD_{\infty}}$,$\mu_{ PSD_{\infty} }$}, %legend style={at={(0.5,-0.6)},anchor=north}, %legend style={font=\tiny}, legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\gamma_{100}}$,$\mu_{\gamma_{100}}$,$\gamma$}, legend pos=south east,% reverse legend ] \pgfplotstableread{NU_KNN_data_singlegraph.dat}{\NUKNNSINGLE} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\NUKNNSINGLE}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\NUKNNSINGLE}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\NUKNNSINGLE}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:nonuniform:knn_singlegraph} \end{subfigure} %legend entries={$\pm \sigma_{\langle\gamma_\infty(h)\rangle_G}$,$\mu_{\langle\gamma_\infty(h)\rangle_G}$,$\gamma$} %%%%%%%% \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure7.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% %ylabel={$PSD_{\infty}(\lambda)$}, %legend entries={$\mu_{ PSD_{\infty}}\pm \sigma_{ PSD_{\infty}}$,$\mu_{ PSD_{\infty} }$}, %legend style={at={(0.5,-0.6)},anchor=north}, %legend style={font=\tiny}, legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\langle\gamma_\infty(h)\rangle_G}$,$\mu_{\langle\gamma_\infty(h)\rangle_G}$,$\gamma$}, legend pos=south east,% reverse legend ] %legend entries={$\mu_{\langle\gamma_\infty(h)\rangle_G}\pm \sigma_{\langle\gamma_\infty(h)\rangle_G}$,$\mu_{\langle\gamma_\infty(h)\rangle_G}$,$\gamma$}, %legend style={at={(0.5,-0.6)},anchor=north} %legend style={font=\small}, %legend pos=south east \pgfplotstableread{U_data_100graphs.dat}{\UDataHundred} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\UDataHundred}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\UDataHundred}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\UDataHundred}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:uniform:noknn_100graphs} \end{subfigure} %% \begin{subfigure}[b]{0.49\columnwidth} \centering \includegraphics[]{IEEE_GLOBALSIP_2018_ASERRANO-figure8.pdf} \begin{comment} \begin{tikzpicture} %\pgfplotstableread{U_data_100graphs.dat}\mytable \begin{axis}[ tiny, width = 5cm,height=3.5cm,% %xlabel={Graph Frequency, $\lambda$}, xlabel={$h$},xlabel near ticks,xlabel shift=-0.1cm,x label style={font=\tiny},% xmin=0, xmax =1, xtick={0,0.2,...,1},% ymin=0, ymax=1.5, ytick={0,0.5,...,1.5},% legend style={nodes={scale=0.5, transform shape},font=\large}, legend entries={$\pm \sigma_{\langle\gamma_\infty(h)\rangle_G}$,$\mu_{\langle\gamma_\infty(h)\rangle_G}$,$\gamma$}, legend pos=south east,% reverse legend ] \pgfplotstableread{NU_data_100graphs.dat}{\NUDataHundred} \addplot[mark = none,color = blue, thick] table[x index=0,y index=2] {\NUDataHundred}; \addplot[mark = none,color = blue, thick,forget plot] table[x index=0,y index=3] {\NUDataHundred}; \addplot[mark = none,color = red, thick] table[x index=0,y index=1] {\NUDataHundred}; \addplot[mark = none,color = black, thick,dashed] table[x index=0,y index=1] {\GTvariogram}; \end{axis} \end{tikzpicture} \end{comment} \vspace{-0.2cm} \caption{} \label{fig:nonuniform:noknn_100graphs} \end{subfigure} \vspace{-0.3cm} \caption{Left column uniform sampling. Right column nonuniform sampling. (\subref{fig:uniform:noknn_singlegraph})-(\subref{fig:nonuniform:noknn_singlegraph}) Statistics of global variogram using 1 fully connected graph and 1000 signal realizations according to $\gamma(h) = 1 - \exp \big(-\frac{h}{0.2}\big)$. (\subref{fig:uniform:knn_singlegraph})-(\subref{fig:nonuniform:knn_singlegraph}) Statistics of global variogram using a single 100-nearest neighbor graph and 1000 signal realizations. There is a shorter support for the global variogram in the uniform sampling scheme since KNN removes most samples across longer distance ranges for uniform structures than it does for the nonuniform structure. (\subref{fig:uniform:noknn_100graphs})-(\subref{fig:nonuniform:noknn_100graphs}) Statistics of global variograms over 100 fully connected graphs each with 1000 signal realizations. }
\caption{Overview of Multi-Scale Vehicle Representation (MSVR) learning for discriminative vehicle re-id at varying spatial resolutions. MSVR learns vehicle re-id sensitive feature representations from image pyramid by an network architecture of multiple branches all of which are optimised concurrently (consensus feedback shown in {\color{red} red}, see Eq.~\eqref{equ:cross-entropy}) subject to the same ID label constraints. Importantly, an inter-scale interaction mechanism is enforced to further enhance the scale-generic feature learning. % {\bf \color{red} TODO for Aytac. (Similar as Yanbei's but try to be different).} }
\caption{ Adaptation from GTA5~\cite{richter2016gta5} to Cityscapes~\cite{cordts2016cityscapes}. We present per-class IoU and mean IoU. ``V'' and ``R'' represent the VGG16-FCN8s and ResNet101 backbones, respectively. ``ST'' and ``AT'' represent two lines of method, \emph{i.e.,} self training- and adversarial learning-based DA. We highlight the best result in each column in \textbf{bold}. To clearly showcase the effect of CLAN on infrequent classes, we highlight these classes in \textcolor{blue}{blue}. \emph{Gain} indicates the mIoU improvement over using the source only. }
\caption{ Adaptation from SYNTHIA~\cite{ros2016synthia} to Cityscapes~\cite{cordts2016cityscapes}. We present per-class IoU and mean IoU for evaluation. CLAN and state-of-the-art domain adaptation methods are compared. For each backbone, the best accuracy is highlighted in \textbf{bold}. To clearly showcase the effect of CLAN on infrequent classes, we highlight these classes in \textcolor{blue}{blue}. \emph{Gain} indicates the mIoU improvement over using the source only. }
\caption{\label{FIG:det-concept}Event browser for simulations using a development version trigger algorithm and Geant4 detector mockup. Color convention (see electronic version): \textcolor{blue}{pion}, \textcolor{magenta}{photon}, \textcolor{green}{electron}(dashed), \textcolor{red}{muon}. This is a side view; the umbrella, inner cube and Si(Li) layer outlines are drawn.}
\caption{(a) Schematic presentation of the assembled setup. \blue{The} Maltese cross \blue{(shown near CCD) is a polar plot of the mean intensity with variable $\theta-\theta_R$} (Eqn.~\ref{es}). A 10-nm-bandwidth filter (see, the actual spectral profile of transmission, $T$, plotted) was inserted to filter out the rest of the white light condenser illumination. Arrows mark \blue{transmission orientation of the} polariser and analyser. (b) Retardance vs voltage of liquid crystal cell (LCC1223T-A; Thorlabs) at two wavelengths \blue{calibrated by manufacturer.}}
\caption{Optical micro-images of white silk \blue{ (Bombyx mori)} through the crossed polarizer-analyser under the white light illumination. Liquid crystal (LC) retarder voltage is marked. Slow-axis of the LC retarder was at $\theta=45^\circ$ and silk fiber was perpendicular to the slow-axis of the LC-retarder.}
\caption{(a) Optical image of a brown silk \blue{(Antheraea pernyi)} fiber at the liquid crystal cell retarder voltage of 15~V. The band pass filter at 635~nm wavelength was used for the white light condenser illumination. The birefringence of silk is determined by the difference of the extraordinary and ordinary refractive indices $\Delta n = n_e -n_o$. The dark region (out off the fiber sample) is where only the LC-retarder is between crossed polariser-analyser, marked as region of interest (ROI): ``Air''. (b) The intensity integrated over selected ``Air'' (see, (a)) vs. the voltage (rms) of the retarder. (c) Digitised retardance of the LC cell $Ret_{exp}=\arcsin\sqrt{T_\theta}$, where $T_\theta = I_\theta/I_0$ (Eqn.~\ref{es}). Temperature of the LC-retarder was set 24.7-24.9$^\circ$C, exposure time 0.85~s. }
\caption{Experimental measurement of the transmission. (a) Integrated transmission intensity vs. retardance $Ret_{exp} = \arcsin\sqrt{T}$ for the marked region (see the inset) without silk fiber (marked square: ``Air''). Inset shows the image of the brown silk \blue{(Antheraea pernyi)} fiber at 15~V with highlighted regions of the birefringence measurement $2\times 2$ pixels ($0.66\times 0.66~\mu$m$^2$); note, the resolution of the used objective lens was $\sim~2~\mu$m. (b) Integrated transmission through the silk fiber (dots) and the best fit (line). The fit function $fit(x)=a\sin^2(\pi[x + b])+o$, where the best fit was achieved for selected \blue{amplitude, phase delay, and offset $a$, $b=\Delta nd$, and $o$, respectively,} and $\Delta n= 2.17\times 10^{-2}$ when $d = 30~\mu$m. }
\caption{Retardance $\Delta nd/\lambda$ map at four different wavelengths selected by interference filters (635, 575, 525, 425~nm) measured through the brown silk \blue{(Antheraea pernyi)} single strand. Since silk is naturally made from two strands, after degumming an asymmetry of the strand is revealed (cross section has a trapezoidal or triangular shape). The average $2\times 2$ pixels was used for numerical processing of the original VGA $480\times 640$ pixels CCD images. The optical resolution can be estimated as \blue{the} radius of Airy disk $w = 0.61\lambda/NA = 0.95~\mu$m for the used $NA = 0.4$ objective lens. The silk fiber was placed on the LC-retarder which had the slow axis orientation parallel to the fiber. The ROI region was used to determine birefringence. }
\caption{Retardance $\Delta nd/\lambda$ vs $1/\lambda$ at four wavelengths (Fig.~\ref{f-image}) \blue{635, 575, 525, 425~nm presented by corresponding color markers}. Each point is average over ROI (see, Fig.~\ref{f-image}). Thickness of the brown silk \blue{(Antheraea pernyi)} fiber was $d \approx 30~\mu$m, which defines the birefringence $\Delta n\approx (1.63\pm 0.05)\times 10^{-2}$. \blue{The linear fit equation $y=p_1x+p_2$, coefficient and 95\% confidence interval are $p_1 = 487.5$ (from 450.6 to 524.5), $p_2=-0.3929$ (from -0.4639 to -0.3219), $R^2=0.9994$.}}
\caption{Wavelength-averaged retardance ($\Delta nd$~[nm]) map (averaged over the four wavelengths; Fig.~\ref{f-image}). Thickness of the brown silk \blue{(Antheraea pernyi)} fiber was $d \approx 30~\mu$m, which defines average $\Delta n = (1.63\pm 0.05)\times 10^{-2}$ at the center of the silk fiber.}
\caption{\label{fig:FigureOfMerit}Examples of the figure of merit, $\Phi_\text{TC}$, vs number of particles per unite area, $m=256$. \full\$\mathrm{SD} = 0$, \broken\$\mathrm{SD} = 0.5$, \dashed\$\mathrm{SD} = 1$. ($a$)~$s=0$; ($b$)~$s=1$, the electrical conductivity was computed along direction of fillers alignment.}
\caption{\label{fig:conds05m512}Examples of the dependencies of the logarithm of the averaged effective electrical conductivity, $\sigma$, on the concentration, $p$, at \fullsquare~$\mathrm{SD} = 0$, \fullcircle~$\mathrm{SD} = 0.1$, \opentriangle~$\mathrm{SD} = 0.5$, \opendiamond~$\mathrm{SD} = 1.0$ and $s = 0.5$, $m = 512$, $L = 32$ ($a$)~along the direction of alignment of the rods, ($b$)~in the perpendicular direction.}
\caption{\label{fig:condSD01s1m1024}Examples of the dependencies of the effective electrical conductivity, $\sigma$, on the concentration, $p$, $\mathrm{SD} = 0.1$, $s = 1$, $m = 1024$, $L = 32$ along the direction of alignment of the rods (\fullsquare) and in the perpendicular direction (\fullcircle).}
\caption{\label{fig:condSD1m1024}Examples of the dependencies of the effective electrical conductivity, $\sigma$, on the concentration, $p$, $\mathrm{SD} = 1$, $s = 1$, $m = 1024$, $L = 32$ along the direction of alignment of the rods (\fullsquare) and in the perpendicular direction (\fullcircle): ($a$)~$s = 1$; ($b$)~$s = 0.8$.}
\caption{\label{fig:pcm512}Examples of the normalized percolation threshold, $p_c/p_{c0}$, versus $\mathrm{SD}$ for three values of the standard deviation $\mathrm{SD}$ at $s$ = 0.5 (\opensquare, \fullsquare), 0.8 (\opencircle, \fullcircle), 1.0 (\opentriangle, \fulltriangle) (closed and open symbols correspond to the direction along alignment of the rods and in the perpendicular direction, correspondingly) and $m$ = 256. Here, $p_{c0}$ corresponds to $\mathrm{SD}$ = 0.}
\caption{\label{fig:IC}The intrinsic conductivity along the direction of alignment of the rods, $[\sigma_\parallel]$, vs the aspect ratio, $k$, of the rods (closed symbols, solid lines) and the intrinsic conductivity in the perpendicular direction, $[\sigma_\perp]$ vs $1/k$ (open symbols, dotted lines) at $\mathrm{SD}$=0 (\opensquare, \fullsquare), 0.1 (\opencircle, \fullcircle), 0.5 (\opentriangle, \fulltriangle), 1 (\opentriangledown, \fulltriangledown) for the case of completely aligned rods ($s = 1$).}
\caption{(Color online) \red{Mean sea surface current in the Gulf of Mexico averaged within $0.3\times0.3$ bin from the 2000-2016 Global Drifter Program trajectory dataset. White lines show the $100$, $1000$ and $3000$ m isobaths. Purple triangles indicate the initial positions of the drifters deployed during the GLAD program. YC and FS represent Yucatan Channel and Florida Straits respectively. Figure is plotted using MATLAB R2014b (http:// www.mathworks.com/) with the M\underline{\hspace{0.2cm}}Map (a mapping package, http://www.eos.ubc.ca/$\sim$rich/map.html).} }
\caption{Example of results obtained (using \emph{MagicHaskeller} as IP core) compared with \emph{FlashFill}. \emph{Output} is the expected output. The first row of each dataset (\emph{id}) is the example given to \emph{FlashFill} and \emph{MagicHaskeller} to learn. {\color{verde}{Green}} and {\color{red}{Red}} colours mean, respectively, correct and incorrect results. The accuracy is $correct\_examples/(total\_examples-n)$, where $n=1$. }
\caption{Additional \aastex\symbols}
\caption{A three node system, with the flow on the edge indicated in blue.\textcolor{red}{$q_k$ needs to be reassigned in the figure right now it looks like going to i}}
\caption{Quantitative comparison on four large-scale datasets. The best two results are shown in \textcolor[rgb]{1,0,0}{red} and \textcolor[rgb]{0,0,1}{blue}, respectively.}
\caption{FER versus $E_{\tb}/N_0$ for TMP and \GL{unquantized} BP decoding for $R = 3/4$ (\protect\tikz[baseline=-0.5ex]{\protect\draw[line width=1,TUMBeamerGreen] (0,0) -- (.5,0)}), $R = 5/6$ (\protect\tikz[baseline=-0.5ex]{\protect\draw[line width=1,TUMBeamerOrange] (0,0) -- (.5,0)}) and $R = 7/8$ (\protect\tikz[baseline=-0.5ex]{\protect\draw[line width=1,TUMBeamerRed] (0,0) -- (.5,0)}). We compare the TMP performance of optimized codes (\ref{plt:tmp_34}, \ref{plt:tmp_56}, \ref{plt:tmp_78}) to their AR4JA counterparts with unquantized \ac{BP} (\ref{plt:full_34}, \ref{plt:full_56}, \ref{plt:full_78}) and TMP decoding (\ref{plt:ar4ja_tmp_34}, \ref{plt:ar4ja_tmp_56}, \ref{plt:ar4ja_tmp_78}).}
\caption{Output hypnogram (a) produced by the proposed SeqSleepNet ($L=20$) for subject 22 of the MASS dataset compared to the ground-truth (b). The errors are marked by the \textcolor{red}{$\times$} symbol. The posterior probability distribution over different sleep stages is shown in (c).}
\caption{Illustration of an \textit{elementary strong collapse}. In the complex on the left, $v$ is dominated by $v'$. The link of $v$ is highlighted in {\color{red} red}. Removing $v$ leads to the complex on the right.}
\caption{Left: $K$ (in grey), Right: $\mathcal{N}(K)$ (in grey) and $\mathcal{N}^2(K)$ (in \blue{blue}). $\mathcal{N}^2(K)$ is isomorphic to a full-subcomplex of $K$ highlighted in \blue{blue} on the left.}
\caption{Schematic view of the conversion process of Belle (\textcolor{Tbluelight}{light blue}) to Belle~II (\textcolor{Tbluedark}{blue}) mDST files using the \texttt{BASF2} modules (\textcolor{Tgraydark}{gray}) provided by the \texttt{b2bii} package and the original Belle software provided by the \texttt{belle\_legacy} library (\textcolor{Tgraydark}{gray}).}
\caption{Matching of the Belle \texttt{PANTHER} Tables (\textcolor{Tblue!50}{light blue}) to the Belle~II \texttt{ROOT} objects (\textcolor{Tbluedark}{blue}) and relations (\textcolor{Torangedark}{orange}).}
\caption[Caption for Quantitative results]{Quantitative results on OTB100, OTB50 and OTB2013. Overlap success (OS) and distance precision (DP) are reported according to the AUC score and the error threshold of 20 pixels (in Figure \ref{fig.overalReslts}), respectively. Top three results are indicated in \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} font. }
\caption{ ImageNet comparison. We report top-1, top-5 accuracy on ImageNet compared with state-of-the-art prior methods. For each DNN architecture, rows are sorted in number of bits.Baseline results were token from PyTorch model zoo. Compared methods: \textcolor{color_joint}{JOINT} \cite{2018arXiv180805779J}, \textcolor{color_pact}{PACT} \cite{choi2018pact}, \textcolor{color_lq}{LQ-Nets} \cite{ZhangYangYeECCV2018}, \textcolor{color_faq}{FAQ} \cite{mckinstry2018disc}}
\caption{{\color{red} Spectrogram (left) and S-method result (right) for Figure \ref{fig:yboth} ($h_\times$) in a logarithmic color scale. We adopt a window width in equation (\ref{eq:hamming}) of $l = 0.0502$ s for the spectrogram and $l = 0.1507$ s for the S-method and frequency width of $\theta_L= 24.9$ [Hz] for the S-method. We here use the smaller window for the spectrogram because $l = 0.1502$ s is too large to secure a sufficient time resolution. \label{fig:spectS}}}
\caption{{\color{red}Slice of the TFRs at $t=0.25$s. The orange solid and green dashed lines indicate the slices of the spectrogram (window size= 0.05 s) and the S-method (window size=0.15 s) in Figure \ref{fig:spectS}, respectively. The blue solid line corresponds to the spectrogram with the window size of 0.15 s.\label{fig:slice}}}
\caption{Mode identification by IF tracking via the S-method. {\color{red} The theoretical prediction of the PNS $g$-mode oscillation, which represents the peak frequency according to \citet{BMuller13}, is shown by the white dashed curve.} \label{fig:smif}}
\caption{\color{red} Comparison of the mode identification by the spectrograms and the S-method. The gray lines are the IF tracked by the spectrogram with different window sizes; light and dark lines corresponding to the window size of 0.15 sec and 0.05 sec, respectively. The color lines are the IF by the S-method, same as Figure \ref{fig:smif}. \label{fig:smif_diff}}
\caption{Stokes parameters in TFRs ($U$ mode) of Figure \ref{fig:yboth}. {\color{red} The left and right upper panels show the spectrogram-type (window size of 0.15sec) and S-type (window size of 0.05sec) Stokes $U$ parameters}, respectively. The bottom panels show {\color{red}$|U|$} from the upper panels on a logarithmic scale. The time and frequency windows are the same as those in Figure \ref{fig:spectS}. \label{fig:compV4}}
\caption{Right: S-method result of the GWs from each shell. Left: Extracted IFs overlaid onto the right panels. The colors are the same as in Figure \ref{fig:smif}. The dashed lines are the aliased modes of A and C for $n=1$. {\color{red} The Nyquist frequency of the simulations is $f_{\mathrm{Ny}} = 500$ Hz.} \label{fig:shell}}
\caption{The value of $h_\im(k_o,\xi)$ (blue curve) as a function of $\xi$ (in units of $4\smash{\sqrt{2}}q_o$) in the wedge in a few representative cases. (a) $k_o = 1.24 + 1.1 i\blue{\in D_1}$ ($V_o=6$, transmission). (b) $k_o = 0.216 + 1.1 i\blue{\in D_2^+}$ ($V_o=2$, \blue{trapping}). (c) $k_o = \blue{0.0479 + 0.95 i\in D_2^-}$ ($V_o=\blue{5}$, \blue{trapping plus wake}). (d) $k_o = 0.214 + 0.5 i\blue{\in D_3}$ ($V_o=8$, \blue{transmission plus wake}). Black dots: $V_o$; red dots: $V_*$ \blue{(trapped soliton); blue dots: $\xi_*$ (wake).} }
\caption{(a) Amplitude difference between the solution shown in Fig.~\ref{f:1}\blue{(a)} and a solution without the corresponding soliton (as in Fig.~\ref{f:4}), demonstrating that in this case the interaction produces no lasting effects on the wedge. %Top left: transmission. %Top right: trapping. %Bottom left: mixed regime. %As in Fig.~\ref{f:4}, %the dashed and solid red lines denote trajectories according to $V_o$ and $V_*$, respectively, %while the blue lines denote the boundary of the wedge. (b) full width at half maximum (vertical axis) of two solitons with the same velocity, one generated by a discrete eigenvalue $k_o\in D_1$ (solid lines) and the other by an eigenvalue $k_o\in D_3$ (dashed lines). The horizontal axis is $V_o$ in units of $4\sqrt2q_o$. Black lines: $k_\im = 0.8$; Blue lines: $k_\im = 0.6$; Red lines: $k_\im = 0.4$; Magenta lines: $k_\im = 0.2$. }
\caption{% Density plots of numerical simulations of Eq.~\eref{e:NLS} with \beginblue initial conditions \endblue consisting of a soliton plus a Gaussian perturbation of the constant background \blue{with $q_o=1$}. The horizontal axis is~\blue{$x$}, the vertical axis is~\blue{$t$}, and the grayscale shows $|q(x,t)|$. Recall that the wedge generated by an initial disturance localized at $x=0$ is confined to the region $|x|<4\sqrt2q_ot$ \cite{biondinilimantzavinos}. \beginblue(a) \endblue $k_o = 3 + 0.5\,i$ (implying $V_o= 12.7$), resulting in a soliton transmission. \beginblue(b) \endblue $k_o = 0.3 + 1.5\,i$ (implying $V_o= 1.63$), resulting in a soliton trapping. }
\caption{% \beginblue(a) \endblue Same as Fig.~\ref{f:1}, but for $k_o = 0.1 + 0.5\,i$ (implying $V_o= 15.5$), resulting in a mixed regime in a soliton transmission is accompanied by the formation of a wake in the wedge. \beginblue(b) \endblue Contour lines of constant soliton velocity in the spectral plane and the domains $D_1$ (gray), $D_2$ \blue{(dark and light blue)} and $D_3$ (orange) of the spectral plane resulting in the \blue{various} outcomes. \blue{(The red curve is determined by the modulation equations in \cite{biondinilimantzavinos}.)} }
\caption{% Density plots of the difference between the amplitude of a\break solution with a soliton and that of a solution without the soliton (i.e., generated by just the localized disturbance), demonstrating the presence of the soliton inside the wedge, the change in the soliton velocity and the generation of the soliton wake. \beginblue(a) \endblue $k_o = 0.1 + 1.02\,i\in D_2$ (implying $V_o = 1.95$), resulting in a soliton trapping. \beginblue(b)~\endblue $k_o = 0.2+0.5\,i\in D_3$ (implying $V_o= 8.42$), resulting in a mixed regime comprising a soliton transmission and a soliton-generated wake. Blue lines: the boundaries $x=\pm 4\sqrt2q_ot$ of the wedge. Dashed red lines and solid red lines: trajectories corresponding to $V_o$ and to $V_*$, respectively. }
\caption{% Effects of the interaction between the soliton and the wedge. \beginblue(a) \endblue Velocity $V_*$ (vertical axis) of the trapped soliton after it enters the wedge as a function of the incident soliton velocity $V_o$. (Black: $k_\im = 1.02$; blue: $k_\im = 1.05$; red: $k_\im = 1.1$; magenta: $k_\im = 1.4$. Also shown for comparison is the dashed gray line $V_* = V_o$.) \beginblue(b) \endblue Velocity $V_*$ of the soliton-generated wake as a function of $V_o$. (Black: $k_\im = 0.8$; blue: $k_\im = 0.6$; red: $k_\im = 0.4$; magenta: $k_\im = 0.2$.) %Right: (transmission): The soliton position shift $\Delta X$ (vertical axis) as a function of the soliton velocity $V_s$. All velocities are in units of $4\sqrt2q_o$. }
\caption{A DeRRT$^*$-based planner starts {\red at the red square} and tries to reach the {\green green square} while escaping from a bug trap. % The {\blue search tree, shown as blue circles}, is mirrored by a sequence model, an HMM or LSTM\@.% When expanding the tree, a free-space sample is drawn, steered toward, and the {\red resulting node, shown as a red circle}, is used to find the closest node in the tree; as in RRT\@.% The sequence model, with state corresponding to that closest node, observes this free-space sample, the path leading to this node, along with {\grey local visual or map features, shown in gray}, and predicts a {\green modified direction, shown in green}, which is then connected to the search tree. % A new state for the sequence model is also predicted and connected. % This process incorporates the bias to explore free space of RRT-based planners with a co-evolving sequence model and observations of the environment.}
\caption{{\it Left panel:} Gas velocity field (shown by the white vectors) in a slice through a galaxy cluster simulation after subtracting the global mean bulk velocity of the cluster. The underlying colour map represents the gas temperature of this relaxed cluster. The black circles show $r_{2500}$, $r_{500}$ and $r_{200}$. Figure taken from \citet{nelson14}, reprinted with permission. {\it Right panel:} The profile of non-thermal pressure fraction, $f_{\rm nth} \equiv P_{\rm nth}/P_{\rm total}$ (where the total pressure, $P_{\rm total}$, is the sum of the thermal pressure, $P_{\rm th}$, and the non-thermal pressure, $P_{\rm nth}$), as a function of the cluster-centric radius scaled with $r_{200m}$ at $z=0$. Modelled and simulated profiles of an early growth sub-sample ($\Gamma_{200m} < 1.8$, red lines) and a late growth sub-sample ($\Gamma_{200m} > 2.7$, \blue lines) are shown. This plot compares the predicted profiles from the analytic model of \citet{shi14} (modeled) with the profiles for a cosmological hydrodynamics simulation from \citet{nelson14b} (simulated), showing good agreement between the two methods. The lines and the shaded regions are the mean profile of the sample and the 16/84 percentiles, respectively. Figure taken from \citet{shi15}, reprinted with permission.}
\caption{Set-based Agreement View Representation where label accuracy is represented in the documents column as: \textcolor{OliveGreen}{\textless{}correct\textgreater{}}, \textcolor{BrickRed}{\{wrong\}} and \textcolor{Blue}{[unknown]}.}
\caption{Reduced non-normalized probability density $\tilde{f}(\tilde{x})$ as a function of the reduced variable $\tilde{x} = x/l$ for $\sigma = 0.4$ and $\sigma = 0.8$. The solid horizontal lines represent the theoretical result \textcolor{blue}{(\ref{f1,a1})} for $\tilde{f}$, and triangle symbols represent the results of numerical simulations of \textcolor{blue}{Eq.\(\ref{eq_X})}. The theoretical values of the probability $a$ ($a=0.4$ for $\sigma = 0.4$ and $a=0.3$ for $\sigma = 0.8$) are also in good agreement with the numerical ones ($a \approx a_{-} \approx a_{+}$).}
\caption{Reduced non-normalized probability density $\tilde{f}(\tilde{x})$ as a function of the reduced variable $\tilde{x} = x/l$ for $\sigma = 1.3$ (a) and $\sigma = 1.7$ (b). The solid lines show theoretical dependencies obtained from \textcolor{blue}{ (\ref{red_f2})} and \textcolor{blue}{(\ref{alpha,beta})}, and the triangle symbols indicate the results obtained by numerical simulations of \textcolor{blue}{Eq.\(\ref{eq_X})}.}
\caption{Reduced non-normalized probability density $\tilde{f}(\tilde{x})$ as a function of the reduced variable $\tilde{x} = x/l$ for $\sigma = 2.3$ (a) and $\sigma = 2.7$ (b). The solid lines represent the theoretical results obtained using \textcolor{blue}{(\ref{red_f3})}, \textcolor{blue}{(\ref{mu})} and \textcolor{blue}{(\ref{nu})}, and the triangle symbols show the numerical results obtained by numerical simulations of \textcolor{blue}{Eq.\(\ref{eq_X})}.}
\caption{Probability $a$ of the extremal values of the bounded process $X_{t}$ as a function of the ratio parameter $\sigma$. The solid lines represent the theoretical results \textcolor{blue}{(\ref{f1,a1})}, \textcolor{blue}{(\ref{a2})} and \textcolor{blue}{(\ref{a3})} for $\sigma \in (0,1)$, $\sigma \in (1,2)$ and $\sigma \in (2,3)$, respectively. The results obtained via numerical simulations of \textcolor{blue}{Eq.\(\ref{eq_X})} are marked by triangle symbols. Inset: probability $a$ vs.\$\sigma$ for large values of $\sigma$. The solid line represents the asymptotic formula $a=1/2\sigma$.}
\caption{Global optimum and our method. When the map is completely known, the optimal trajectory computed using the approach of \cite{liu2018searchBasedSE3} is shown in blue (\tikzrectangle[black,fill=MyblueLight]{10pt}). The red trajectory (\tikzrectangle[black,fill=Myred]{10pt}) is the solution found by our method, where the world is not known and it is being discovered as the UAV flies forward. The grid is $1$m $\times 1$m, and the sensing range is 10 m.}
\caption{Priority queue $\mathcal{K}$. Given $JPS$ (\tikzrectangle[black,fill=Myblue]{10pt}), the priority queue $\mathcal{K}$ returned by $SamplePoints(\bb{B'})$ contains points in this order: First samples along the spherical arc $\bb{B'} \rightarrow \bb{P}_{\bb{q}_{3}}$. Then along the arc from $\bb{P}_{\bb{q}_{3}}$ to the next projection, and so on. After that, several samples are taken from concentric circumferences to $\bb{B'}$. }
\caption{Instantaneous sensing data and occupancy grid pipeline. Sensing data from the depth sensor (\tikzrectangle[black,fill=MyorangeDarker]{10pt}) is received at $f_{sensor}$. New point clouds are fused into the Occupancy Grid (\tikzrectangle[black,fill=MygreenDark]{10pt}). Collision check at $t_{1}$ is done using an array of \textit{k}-d trees that contains the \textit{k}-d tree of the last map fused (\tikzrectangle[black,fill=MyorangeDark]{10pt}), and the \textit{k}-d trees of some of the last point clouds received that are not included in the map (\tikzrectangle[black,fill=MyblueLight]{10pt}).}
\caption{Office simulation. The UAV must fly from $\bb{A}$ to $\bb{G_{term}}$ in an office environment. The optimal trajectory is shown in blue (\tikzrectangle[black,fill=MyblueLight]{10pt}). The red trajectory (\tikzrectangle[black,fill=Myred]{10pt}) is the solution found by our method.}
\caption{Generalized group testing with inhibitors paradigm encompasses two procedures: encoding and decoding. For the theoretical model, the objective of encoding procedure is to create a pooling design (i.e., a measurement matrix) and then do tests on it to get outcomes. From the outcomes, the items are classified via the decoding procedure. There are nine steps in the practical model while steps 4 to 6 merge into one step in the theoretical model. In the practical model, a set of items might consist of \textcolor{red}{defective} items ($\color{red} \bullet$), \textcolor{violet}{inhibitors} ($\color{violet} \bullet$), \textcolor{blue}{hybrid} items ($\color{blue} \bullet$), and negative items ($\bullet$). Data gathered from tests must be converted into binary outcomes before proceeding the decoding procedure to classify the items.}
\caption{The spectral density of ER random networks and scale-free networks for different average degrees and $N=1000$. {\color{black} $\vartriangle$}, {\color{blue} $\square$}, {\color{red} $\circ$} and {\color{green} $\ast$} represent the data points of density distribution for $\langle k \rangle$=2, 4, 6 and 8, respectively. Reproduced figure (Figure 2) from A. Yadav and S. Jalan, Chaos 25, 043110 (2015), with the permission of AIP Publishing. DOI: \href{https://aip.scitation.org/doi/10.1063/1.4917286}{10.1063/1.4917286}.}
\caption{The evolution of the transverse wave for a range of dimensionless sphere radii $\drs=R_s/\ell_{*}$ in the range $0.34\leq\drs\leq4.29$ (a) Measurements of the radial position of a capillary wave on a bare interface ($\times$) gives $r_{m} \sim t^{2/3}$, different from the behaviour with an ultra-thin PS sheet (coloured points). Inset: wave propagation depends on $\Rf$ (= 5.51\,mm for\textcolor{marker12}{$\lhd$} and 13.30\,mm for\textcolor{marker22}{$\circ$}). (b) Rescaling the data using the horizontal length scale $\ell_{*} = (\glv \Rf/\rho V^2)^{1/2}$ and associated timescale $t_{*} = \ell_{*}/V$ shows that $r_{m} \sim (\ell_{*}Vt)^{1/2}$, as in (\ref{eq:wavescaling}). The associated theoretical predictions (including prefactors) are shown for $\alpha = 0.34$ (solid line) and $\alpha=4.29$ (dashed line). (c) The experimental prefactor $\beta=r_m/\sqrt{\lstar Vt}$ in \eqref{eq:wavescaling} varies with $\drs$ in a similar way to that predicted theoretically~\cite{prfluids} (solid black line) at sufficiently early times (as defined in the SI). %Note: The data markers are colour coded as a function of $\drs$, in accord with the values given in (c), and experimental parameters are given in the SI. Throughout, data markers are colour coded according to the value of $\alpha$ in (c); experimental parameters are given in the SI. }
\caption{Order-two convergence for the travelling wave solution of the extended Burgers' equation outlined in section \ref{sec:P}. The plots correspond to the conventional solution (\marksymbol{o}{blue}) and the collective solution (\marksymbol{triangle}{red}) and an order-two reference line (\RefLine). The error is calculated after 512 timesteps, with $L = 8$, $\Delta t = 2^{-14}$ and $\Delta x = L/2^{k}$ for $k=1,2,3$ and $4$. }
\caption{(a) Cross correlation function $g^{(2)}(0)$ for two particles as a function of relative brightness, $\alpha$. The maximum value of $g^{(2)}(0) = 0.5$ is achieved for equal brightness particles. (b) Overlapped contour plots of $g^{(1)}(r_1,r_2,\alpha)$ ({\color{red} red}) and $g^{(2)}(\uptau=0,r_1,r_2,\alpha)$ ({\color{blue} blue}) for $\alpha=1$, and (c) $\alpha=0.5$ where radial distance is measured in units of the standard deviation of the point spread function of the illumination/collection optics. By comparing both the $g^{(1)}$ and $g^{(2)}$ values, it is possible to determine more information about the particles' positions than is possible using intensity alone. When $\alpha\neq 1$ the symmetry between the contours for $r_1$ and $r_2$ is broken. %is to shift it away from $(r_1,r_2)=(0,0)$. }
\caption{Quantitative comparisons of state-of-the-art methods. \textcolor[rgb]{1.00,0.00,0.00}{Red} text indicates the best performance and \textcolor[rgb]{0.00,0.00,1.00}{blue} italics text indicates the second best performance. We use results from LapSRN to do comparation, and attention that Layers in the table include convolution and deconvolution.}
\caption{\label{fig:wide}(color online) Temperature dependence of $\kappa$ in the V$_{1-x}$Ti$_x$ alloys below the $T_C$ in the superconducting (zero magnetic field, open symbols) and normal states (8~T, closed symbols). The $\circ$ and $\bullet$ represent the experimental data points, {\color{red} $\vartriangle$} and {\color{red} $\blacktriangle$} represent the $\kappa_e$, {\color{blue} $\square$} and {\color{blue}$\blacksquare$} represent the $\kappa_l$. The solid (black) line is the fit to the $\kappa_n$ and the dash-dotted (blue) line is the fit to the $\kappa_{ls}$. In the superconducting state of vanadium, the heat is carried mainly by the electrons. The $\kappa$($T$) in the superconducting state of the V$_{1-x}$Ti$_x$ alloys with $x >$ 0.1 increases when the temperature is decreased below $T_C$ indicating that the phonons are the major carriers of heat in the superconducting state.}
\caption{(a) Reduction in $\phimax/\phimax(t=0)$ (\textcolor{blue}{$\blacktriangleup$}) and $\phimin/\phimax(t=0)$ (\textcolor{red}{$\blacktriangledown$}) with collisional age, for $\tau=50$, and $\Ma=1.15$. For reference, the $\propto\sqrt{\nus t}$ dependence is indicated by the dashed line. (b) The $\Ma$ and $\tau$ dependence of $1-\phimin(t)/\phimin(t=0)$ at the collisional age $\nus t=0.1$. }
\caption{ Three visualizations of an artificially generated dataset with four different classes (\classred, \classgreen, \classblue, \classorange). (a) demonstrates color blending, which uses binned color and saturation to encode class labels and their intensity. (b) demonstrates weaving with bin-internal normalization, resulting in fully filled bins with the number of colored fragments encoding bin-relative class intensities. (c) demonstrates hatching, which uses both angle and color to encode class labels. }
\caption{(Color online) ($a$) Little-Parks oscillations for different values of EXSO parameter $\lambda(\varphi)=\pm\lambda_{so}\cos(\varphi-\alpha)$ for $|\varphi|\leq \pi/2$: $\alpha=0$ -- solid; $\alpha=\pi$ -- dashed. The numbers near the curves denote the corresponding values of the spin-orbit constant $\lambda_{so}$. ($b$) Dependence of the shifts of the Little-Parks oscillations $\Delta\phi$ ({\color{red} $\blacktriangle$}) and the maximal transition temperature $\tau_{max}$ ({\color{blue} $-\blacksquare-$}) (see the panel ($a$) in confusion) on the angle $\alpha$ for $\lambda_{so}=1$. Solid red line shows the dependence described by the Eq.~(\ref{FluxShift}).}
\caption{A LAMBADA example where the final word ``julie'' (with reference chain in brackets) is the answer, $y$, to be predicted from the preceding context $x$. A system must know the two speakers and the current dialogue turn, simple context matching is not sufficient. Here, our model's predictions \textcolor{red}{before} and \textcolor{blue}{after} adding multi-task objective are shown. %Here, recognizing that swartz is the last speaker helps to disambiguate the correct answer. }
\caption{\label{fig:DS_optimization} \textbf{Improving DS device geometry.} \textbf{(A)} Device structure from Ref. [\textcolor{blue}{17}]. \textbf{(B)} Improved device structure (DS$_{\textrm {imp}}$) to reduce edge roughness for reflected electrons from second junction. Ideally, the local gate should be restricted within the triangle enclosed by green dash line to make the region free from edges and electrons can be redirected to the other source by second junction shown by white dash lines. However, it is impossible to maintain smooth potential at the corners. Therefore, the left and bottom sides are kept extended ($L_{ext}$ = 100 nm). \textbf{(C)} Comparison of on-off ratio of DS vs. DS$_{\textrm{imp}}$. DS$_{\textrm{imp}}$ shows less sensitivity to edge roughness.}
\caption{Correct label (\textcolor{green}{green}), human guess (\textcolor{red}{red}) and DeepGeo guess (\textcolor{blue}{blue}) displayed for a few human test cases.}
\caption{(a): Two-temperature model of a metal-superconducting film on an insulating substrate. Only a fraction of the incident optical power $P_{in}$ is absorbed by the film as determined by the coupling efficiency of the apparatus $P_{abs}=\eta_{couple}P_{in}$. The electron subsystem under illumination can be described by the {\bblue equilibrium} Fermi-Dirac distribution with an effective electron temperature $T_e$ exceeding the phonon temperature in NbN $T_{ph}$ and the bath $T_b$ temperature. Further, it is assumed that the $T_e$ is the only parameter that controls the resistivity of the superconducting film. $Z_{eph}$ and $\bblue Z_{esc}$ are the thermal resistances, which determine the heat exchange of the electrons with the acoustic phonons in NbN and of the phonons in NbN with the substrate, respectively. (b): Measured SSPD detection efficiency $D_E$ at $1.55\,\mu$m wavelength at $T_b=1.7$\,K vs the bias current in units of the critical current. The saturation of the bias current dependence signals the regime of unit internal efficiency, which enables us to calibrate the coupling losses as$D_E=\eta_{couple}$. \textcolor{black}{Inset: The sketch of the experimental setup. S denotes the sample, LD -- the laser diode, PC -- the polarization controller, PM -- the power meter and ATT -- the attenuator.}}
\caption{\bblue The $T$-dependencies of the sample resistance in equilibrium (a) and of the measured voltage response under optical radiation (b) for a set of magnetic fields (see the legend). In panel (b) the laser wavelength of 1.55 $\mu$m, the input laser power of $P_{in}=$320\,nW and the DC bias current of$I=100$\,nA were used. Inset: the$T$-dependence of the upper critical field $B_{c2}(T)$ used to evaluate the density of states in NbN. \textcolor{black}{The $T$-dependencies in panels (a) and (b) were taken by measuring the signals during a warm-up at discrete temperature values (15-20 points per curve). At each point the temperature was stabilized for at least 10 minutes with the accuracy of about $\pm$1\,mK.}}
\caption{Temperature dependence of the thermal resistance $Z$ for NbN SSPD at resistive transition. {\bblue The experimental datasets are obtained from the measured power-response of our SSPD in different magnetic fields, as explained in section~\ref{resistivity}. The dashed guide line represents the power-law dependence $Z\propto T^{-3}$ observed. The $B=0$ critical temperature of the resistive transition is marked by an arrow. Inset: the ratio of the electron (normal-state) and phonon thermal capacitances as a function of temperature. The solid line is the upper bound imposed by our experiment, as discussed in the text. The symbol is the lower bound, which follows from the theory of Ref.~\cite{vodolazov} for the parameters close to our SSPD detector.}}
\caption{Normal and reduction convolutional cell architectures found by \ourmethod{} applied to NASNet micro search space. The inputs (\textcolor{mygreen}{green}) are from previous cells' output (or input image). The output (\textcolor{myyellow}{yellow}) is the results of a concatenation operation across all resulting branches. Each edge (line with arrow) indicates an operation with operation name annotated above the line.}
\caption{\textbf{Crossover Example:} A crossover (denoted by $\otimes$) of a VGG-like structure with a DenseNet-like structure may result in a ResNet-like network. In the figure, {\color{red}{red}} and {\color{blue}{blue}} denotes connections that are unique to VGG and DenseNet respectively, and black shows the connections that are common to both parents. All black bits are retained in the final child encoding, and only the bits that are not common between the parents can potentially be selected at random from one of the parent.}
\caption{Parameter sensitivity analysis of \(\lambda\) on the OTB-2013 dataset. %The results are presented in terms of DP and \(\text{OS}_\text{AUC}\). {\color{Red}{Red}}: best. {\color{Blue}Blue}: second best.}
\caption{Evaluation of baseline improvements using attentive features and attentive classifiers on the OTB-2013 dataset. %The results are presented in terms of DP and \(\text{OS}_\text{AUC}\). {\color{Red}{Red}}: best. {\color{Blue}Blue}: second best.}
\caption{Comparisons with the state-of-the-art trackers on the OTB-2013 and OTB-2015 datasets. Our tracker performs favorably against existing trackers in center location error (CLE), and the overlap success rate at a threshold of 0.5 IoU (\(\text{OS}_\text{0.5}\)). {\color{Red}{Red}}: best. {\color{Blue}Blue}: second best.}
\caption{Comparisons with the state-of-the-art trackers on the VOT-2016 dataset. The results are presented in terms of expected average overlap (EAO), accuracy rank (Ar) and robustness rank (Rr). {\color{Red}{Red}}: best. {\color{Blue}Blue}: second best.}
\caption{ Intel Xeon Scalable experiments for various problem sizes. Setups left of the dotted line use no Intel Optane technology as everything fits into the DRAM, so we switch if off. % Setups right of the dotted line use Optane. Top: Time per degree of freedom update. Bottom: Total energy usage. Each setup is, as long as it fits into the memory, computed \textcolor{textR2}{multiple times (from left to right): We start from a base grid of $27 \times 27 \times 27$ (circles) and test six polynomial orders $p \in \{4,5,6,7,8,9\}$ with $p \leq 6$ denoted through empty symbols. We then add one level of adaptivity ($\Delta \ell =1$, squares). We next rerun the regular grid experiment with a base grid of $81 \times 81 \times 81$ (circles) and one level of adaptivity (squares). A base grid of $27 \times 27 \times 27$ with two levels of adaptivity (triangles) already requires Intel Optane technology unless we choose $p=3$. } \label{figure:code-measurements:optane} % \vspace{-0.5cm} }
\caption{ Energy consumption per degree of freedom on Broadwell (24 cores) and Xeon Scalable (36 cores) for a typical run with a regular and a dynamic grid. The total energy \textcolor{textR2}{plus the energy spent on the memory are given.} % old: =-no-vec, vectorised=AVX2, 100 timesteps, fused, % new: =-no-vec -no-simd, generic, vectorised=AVX2, 50 timesteps, fused, \label{table:code-measurements:energy} }
\caption{ Cut through the solution of the LOH.1 below the surface. Waves propagate from this point but benchmark. A point source induces an ``earthquake'' just yield complicated patterns as the cubic domain contains two layers of different material~\citep{SPICE:06:LOH1}. \textcolor{textR2}{ Regular grid visualization though the experiments run with both regular and dynamically adaptive meshes. } \label{fig:loh1}}
\caption{ \textcolor{textR2}{ Vertical inhomogeneity of the data access cost, and thus speed, arises from multiple cache levels and different cache coherence strategies (inclusive vs.~non-inclusive). % With Intel\textregistered\ Optane\texttrademark's memory mode, With \textcolor{textR3}{the Intel\textregistered\Optane\texttrademark\technology}, main memory \textcolor{textR3}{effectively} becomes a fourth cache and an additional memory layer is added at the bottom. Horizontal inhomogeneity arises from the fact that memory is logically shared yet physically distributed. Further diversity stems from the fact that a core hosts multiple (hyper-)threads which in turn might accommodate multiple logical threads. A third diversity dimension is introduced by the opportunity to calibrate the cores' frequency. } \label{figure:introduction:diversity} \vspace{-0.5cm} }
\caption{ % The Intel Memory Drive Technology (IMDT) operation mode for Optane. % \textcolor{textR1}{ % Intel's second generation Intel Optane DC Persistent Memory eliminates the software IMDT layer and % integrates the large-scale memory directly into the DIMM slots. % } \textcolor{textR2}{ Intel\textregistered\Optane\texttrademark\SSD DC P4800X Series with Intel\textregistered\Memory Drive Technology operation mode. It relies on software IP for memory management. Newer products such as Intel\textregistered\Optane\texttrademark\DC Persistent Memory%, as used for our experiments, integrate are \textcolor{textR3}{promised to integrate} on DIMM form-factor and also \textcolor{textR3}{to} introduce a memory mode fully managed by the CPU without extra software\textcolor{textR3}{; and hence smaller cost penalty}. % The %recommended % \textcolor{textR3}{generic optimization idiom hiding % latency through pipelining however remains valid.} % for both choices available in 2019. } \label{figure:introduction:optane-imdt} \vspace{-0.5cm} }
\caption{Left: Possible birth years and death indicators for 2 children with mother's age $m_{surv}=18$, assuming 1 death. Fertility $f(m)=0$ for $m < 15$. Right: Possible death ages for a child born to a mother of age $m_b=15$. The {\color{red}$\times$}'s represent a death. Here, $q(a)$ is the probability of death by age $a$. A color version of this figure can be found in the electronic version of the article.}
\caption{\small The performance of deep models on valid augmented set. Metric scores are represented by different colors: \textcolor{red}{sAUC}, \textcolor{green}{CC}, \textcolor{blue}{NSS}, \textcolor{cyan}{SIM}, \textcolor{magenta}{AUC-Borji}, and \textcolor{black}{KL}.}
\caption{The overall framework of the proposed method. The tracker can be divided into four parts marked in color: a) \textcolor[rgb]{1.00,0.00,0.00}{Filter learning}; b) \textcolor[rgb]{0.00,1.00,0.00}{Primal detection}; c) \textcolor[rgb]{0.00,0.07,1.00}{Re-detection module}; d) \textcolor[rgb]{1.00,0.50,0.00}{Scale estimation and adaptive update}.}
\caption{The highest correlation coefficients across SMA-EPA pairs for the sample of 33 vendors.\iffalse \textcolor{red}{Transpose the table. Post, Repost and Comment on the side and Search, Clickthrough and Orders on the top. Keep this order precisely }\fi}
\caption{Precision statistics for Random Forest 5-q classifiers for the next-day volumes of each EPA type: Order, Clickthrough and Search. Results are aggregated per five categories. \iffalse\textcolor{red}{import correct statistics}\fi}
\caption{\label{tab:4} Example translations where subjective-case pronouns in brackets are dropped in original input but labeled by DP generator. We italicize some {\em \color{blue} mis-translated} errors and highlight the {\bf \color{red} correct} ones in bold.}
\caption{ {\bf{X-ray diffraction analysis of the septuple/quintuple heterostructures}} formed in \BiTe\and\BiSe\upon Mn doping as a function of Mn concentration ranging from 0 to 11\%. The measured diffraction spectra (red and blue lines in (a,b)) are fitted using a random stacking paracrystal model consisting of a statistically varying alternation of Bi$_2X_3$ quintuple and Bi$_2$Mn$X_4$ septuple layers as described in the Supplementary Information, providing an excellent fit (black lines) of the experimental data for both the telluride and selenide system. The average number of quintuples $\left< N_{\rm QL} \right>$ between subsequent septuples and the root mean square (RMS) width of the random distribution derived from the fit is plotted in (c) versus Mn content (open, respectively, full symbols). A smaller average distance $\left<N_{\rm QL} \right>$, i.e., higher concentrations of septuples, is found for \BiTe\as compared to\BiSe . The number of available Mn sites in the center of the septuple layers relative to the total number $n_{\rm tot}$ of (Bi and Mn) atoms is shown in (d) versus nominal Mn content. The number expected for unity occupancy is indicated by the dashed line. Experimental points below the line indicate that a significant fraction of Mn atoms resides in other lattice sites. This applies to \BiSe\but not to\BiTe. }
\caption{Simulation of six complex organic molecules in a 5$^{\prime\prime}$ source at $T_{\rm{ex}}$~=~100~K, $\Delta V$~=~3~km~s$^{-1}$, and $N_T$~=~5~$\times$~10$^{13}$~cm~$^{-2}$. The simulation assumes the source completely fills the beam. Glycine transitions have been shown in \textbf{\textcolor{red}{bold red}} for emphasis. The approximate noise level of a 10 hour ngVLA integration with 0.6~km~s$^{-1}$ resolution is shown as a dashed \textcolor{blue}{blue} line.}
\caption{{\bf Vertical extent of zombie turbulence.} Root mean square (rms) of relative vorticity and fractional potential temperature anomaly (averaged over $x\!-\!y$ planes) as function of height $z$: (a) Run\_Isothermal after 1792.7 orbits, (b) Run\_Temp\_Step after 2444.6 orbits (turbulent burst phase, shown in red) and 2580.4 orbits (zonal flow phase, shown in black), (c) Run\_Brunt\_Step after 2757.2 orbits (turbulent burst phase, shown in red) and 2797.7 orbits (zonal flow phase, shown in black).}
\caption{Main idea: words (\textcolor{red}{red font}) of test set that hardly appear in training set's common text (bottom-left) are likely to predict named entities. Such words include two kinds: the first kind (\eg `Boston') appear in training set's named entities (top-left) while the second kind (\eg `Reuters') do not. Solid arrow denotes appearing while dashed arrow denotes hardly appearing.}
\caption{Quantitative comparison with state-of-the-art methods, using maximum and average F-measure (larger is better), AUC (larger is better) and MAE (smaller is better). The best three results are colored \color[HTML]{FE0000}\textbf{red}\color{black}, \color[HTML]{3166FF}\textbf{blue}\color{black}, and \color[HTML]{32CB00}\textbf{green}\color{black}, respectively.}
\caption{Instability charts of transverse ({\bf \color{c4} red}) and longitudinal ({\bf \color{c1} blue}) modes. The dashed lines represent different values of $ \kappa $, and resonance bands above $ \kappa = e/ \lambda $ correspond to relativistic production. In the inset we show the $ e / \lambda \ll 1$ limit in which the enhancement of longitudinal mode production over transverse modes is seen explicitly. }
\caption{Viable parameter space for parametric resonance production of vectors in the limit $ \lambda \gg e $. We fix the initial condensate amplitude $ \phi _0 $ such that the dark sector saturates the DM relic abundance $\Omega_X + \Omega_\phi \simeq \Omega_\text{DM}$. Shown are constraints from coldness ({\bf \color{c8} yellow}), cosmic strings ({\bf \color{c4}red}), isocurvature ({\bf \color{c9} purple}), late time dark Higgs decays ({\bf \color{c6} brown}) and sufficiently long PR ({\bf \color{c1} blue}) as described in the text. {\bf Left}: At every point in parameter space we fix $m_\phi \simeq 100 m_X $. We do not incorporate any additional interactions, and the dark Higgs makes up nearly all the DM, i.e. $\Omega_X \simeq 10^{-2} \Omega_\phi$. {\bf Right}: Vectors make up all of the DM, and the dark Higgs is eliminated at late times. At every point in parameter space we fix $ m_\phi = 10 ~{\rm GeV} $ and show the corresponding constraint from thermalization requirements ({\bf \color{c3} green}) as described in Appendix~\ref{sec:appendix}\,.}
\caption{Viable parameter space for parametric resonance production of vectors in the limit $ e \gg \lambda $. We fix the initial condensate amplitude $ \phi _0 $ such that the dark sector saturates the DM relic abundance $\Omega_X + \Omega_\phi \simeq \Omega_\text{DM}$. Shown are constraints from coldness ({\bf \color{c8} yellow}), cosmic strings ({\bf \color{c4}red}), and sufficiently long PR ({\bf \color{c1} blue}) as described in the text. Both plots have the vector making up nearly all the DM. {\bf Left}: At every point in parameter space we fix $m_X \simeq 10 m_X$ and thus $\Omega_\phi \simeq 10^{-1} \Omega_X$. {\bf Right}: At every point in parameter space we fix $m_X \simeq 1000 m_\phi $ and thus $\Omega_\phi \simeq 10^{-3} \Omega_X$.}
\caption{Handshakes in TLS~1.2 with and without session resumption, highlighting \colorbox{mygray}{encrypted} data.}
\caption{Handshakes in TLS~1.3 with and without session resumption, highlighting \colorbox{mygray}{encrypted} data.}
\caption{\underline{Words} indicate diversity compared with the references, while \red{red} words denote translation improvement.}
\caption{ \underline{Words} indicate diversity compared with the references, while \red{red} words denote translation improvement.}
\caption{An example of weighted \acs{semd} for $N=4$. The length of each segment is $L=Nb$, where $b$ is the length of a block. The segments $n$ and $(n+1)$ are overlapped by the length of $(N-1)b$ and each block is shared by four segments. For example, the block marked with \textcolor{red}{$\odot$} is shared by segments 1, 2, 3 and 4. The \acs{emd} process decomposes each segment into the corresponding \acsp{imf}, which will then be multiplied by a weighting function. The weighting function is shown in segments 7 and 8. Then, the \acsp{imf} of all the segments in a shared block are averaged out with the weighting function. For the block \textcolor{red}{$\odot$}, the \acsp{imf} in the last block of the segment 1, in the second last of 2, in the second of 3 and in the first of 4 are weighted-averaged. When the weighting function is a nonzero constant function, the weighted \acs{semd} reduces to the \acs{semd}.}
\caption{ Some examples from the two-class \textit{Freezer} data set. The two classes (\textcolor{blue}{blue}/fine lines vs. \textcolor{red_boston}{red}/bold lines) represent the power demand of the fridge freezers sitting in different locations of the house. The classes are difficult to tell apart globally but they differ locally. }
\caption{Task 1: Mean Absolute Error and Task Completion Time. The \textcolor{blue}{blue} bars show $95\%$ confidence intervals. (A). $Absolute\; error$ = $\left| participant's\; answer - ground\; truth \right|$. (B). Colormaps labeled with the same \textcolor{cyan}{cyan} letter belong to the same group in the post-hoc analysis.} \end{figure} We collected 2496 data points with 24 participants for the two ensemble tasks, or 1728 and 768 for the \textit{ensemble average} and \textit{ensemble set} tasks accordingly. To summarize, the first hypothesis (H1 on rainbow) is partially supported. H2 on multihue, H3 on gray, and H5 on colorfulness are supported. We find no evidence to support H4 on resolution. %Before a statistical analysis was conducted, we used a quantile-quantile (QQ) plot, to test the normality in each of the color and task conditions. We removed outliers if they lay more than three standard deviations from the mean of each experimental condition. In this study, we removed 3 outliers from the data points (2 in TRACING task and 1 in BUNDLE SIZE task). \begin{figure}[!t] \centering \includegraphics[width=0.85\linewidth]{./figures1/t1accuracy.png}% \caption{Colormap Accuracy} \label{fig:t1accuracy} %\end{subfigure} \end{figure} \subsection{Overview of Analysis Approaches and Summary Statistics} Results were analyzed by tasks; Table~\ref{table:exp1:stat} shows the statistical analysis of accuracy and task completion time measured using the following statistical approaches. For both tasks, we examine the main effect of colormap on error and task completion time using the SAS GLM procedure. A post-hoc analysis using the Tukey Studentized Range test (HSD) is performed when we observe significant main effects. %We also further analyze interesting findings when we see them. Task 1 performance is analyzed using several methods. Task completion time is converted to $log_{10}$-based to obtain a close-to-normal distribution. We compute \textit{error} by the distance from the participants' answers to the ground truth and use the formula $error$ = $\log _2 \left|{participant's\; answer} - {ground\; truth} \right|+ 8 $, following Cleveland and McGill~\cite{cleveland1984graphical}. We explore the accuracy of these ensemble colormaps using two additional measurements. \begin{itemize} \item Accuracy. Accuracy is percentage of correct answers. We threshold the error to measure whether an answer is correct. We used $\delta = \left|participant's \;answer - ground\; truth\right|$ and threshold $\delta$ to 0.01-0.04 with step size 0.01. An answer is considered correct when it falls in $\delta$. %\item \textbf{Viewer-specificity.} We further measures whether or not individuals would perform %better using colormaps through correlation analysis. \item Directional Bias. We compute whether or not the colormaps bias observers towards values larger or smaller than ground truth. \end{itemize} The accuracy data in Task 2 are binary and are analyzed using logistic regression and reported using the $p$ value from the Wald $\chi ^2$ test. When the $p$ value is less than $0.05$, variable levels with $95\%$ confidence interval of pairwise difference of odds ratios not overlapping are considered significantly different. The $\chi ^2$ test with the ``$freq$'' procedure is used to examine whether or not there is a significant correlation between the main effect (the colormap or participant) and accuracy. We measure effect sizes using Cohen's $d$ for time and task type I error and Cramer's $V$ for correctness to understand the practical significance~\cite{cohen1988statistical}. We used Cohen’s benchmarks for ``small''(0.07-0.21), ``medium'' (0.21-0.35), and ``large'' ($>0.35$) effects. \begin{figure}[!t] \centering \centering \includegraphics[width=0.98\linewidth]{./figures1/t1colorBias2.png}% \caption{Directional Biases by Colormap. More participants tend to overshoot (report larger than the ground truth) when using isoluminant rainbow. Using the diverging colormap, more participants underestimated the ensemble average. Gray, extended-blackbody, and coolwarm had the minimum directional biases.} \label{fig:t1bias} \end{figure} \begin{figure}[!tp] \centering \centering \includegraphics[width=\linewidth]{./figures1/biasTowardsHigher2.png}% \caption{ Directional Biases by Colormap and Bin. More than $50\%$ larger-than-ground-truth answers appeared in \textit{all} 12 bins for isoluminant rainbow. } \label{fig:biasTowardsHigher} \end{figure} \subsection{Task 1 Ensemble Average Results} For task type 1, ensemble average, colormaps was not a significant main effect on error (Table~\ref{table:exp1:stat} and Fig.~\ref{fig:t1error}). A general trend was that extended blackbody had the least error and gray had the most. Colormap and participant are significant main effects on time. (Table~\ref{table:exp1:stat} and Figure~\ref{fig:t1time}). The post-hoc analysis suggests three Tukey groups: (gray), (blackbody, isoluminant-rainbow, extended-blackbody, and coolwarm), and (blackbody, diverging, extended-blackbody, and coolwarm). The extended-blackbody and coolwarm maps led to the longest task completion time and the gray, though efficient, had the highest error. \subsection{Task 1 Color Sensitivity and Directional Bias} We compute the colormap sensitivity by measuring the percentage of correct answers or accuracy (Fig.~\ref{fig:t1accuracy}). We first compute the mean absolute error. Fig.~\ref{fig:t1accuracy} showed that gray had on average the lowest accuracy among all colormaps. %\begin{figure}[!t] %\centering %\begin{subfigure}{0.32\linewidth} % \centering %\includegraphics[width=0.95\linewidth]{./figures/Charts/t1sensitivity.png}% %\caption{Colormap Sensivity} %\label{fig:t1sensitivity} %\end{subfigure} %\end{figure} Directional bias measures if observers consistently choose larger or smaller values than the ground truth using a colormap. We found that more answers using isoluminant rainbow were biased towards higher values, while the diverging color slightly towards lower answers (Fig.~\ref{fig:t1bias}). All other colormaps of blackbody, extended-blackbody, and coolwarm showed about even distributions between higher and lower participants' answers. We further analyzed the bias distribution in the 12 bins (Fig.~\ref{fig:biasTowardsHigher}). We found that more than $50\%$ of the answers overshoot (selected larger than ground-truth) when using isolumiant-rainbow in \textit{all} bins. Correlations between the data variance and colormap absolute error show that these two variables are statistically significantly correlated for all other maps except the isoluminant-rainbow. This result may indicate that the ensemble behaviors of isoluminant-rainbow might not be as predicable, despite its accuracy for ensemble average is comparable to other colormaps. %Since it is possible that visualization can result in ambiguity and uncertainty inherent to the observer %related to perceptual and cognitive abilities, we further measure whether the rank order of %the color ensembles are participants-dependent. %XXXXXX %\todo[inline, color=green!40]{TODO} \subsection{Task 2 Ensemble Spherical Colormap Results} \begin{figure}[!tp] \centering \subfigure[Absolute Accuracy]{ \includegraphics[width=0.6\linewidth]{./figures1/t2absoluteAccuracy.png}% \label{fig:t2accuracy}} \subfigure[Absolute Task Completion Time]{ \includegraphics[width=0.58\linewidth]{./figures1/t2absoluteTime.png}% \label{fig:t2time}} \label{fig:t2results} \caption{Task 2: Mean Time and Accuracy. The color schemes connected by the \textcolor{orange}{orange} line are significantly different.} \end{figure} The second row in Table~\ref{table:exp1:stat} shows the statistical results. Fig.~\ref{fig:t2accuracy} shows mean accuracy (percentage correct answers) and time and $95\%$ confidence intervals from the mean. Colormap had a significant main effect on accuracy but not on task completion time. H4 is not supported. The Boy's surface embedding and the absolute embedding lead to most accurate answers for following tracts, followed by eigenmap. Boy's surface also shortened task completion time. This task does not require participants to utilize symmetry. The Boy's surface method was more accurate and also fast (Figs.{~\ref{fig:t2accuracy}} and{~\ref{fig:t2time}}). It is also noticeable that the Boy's surface and absolute maps improved accuracy by $15\%$ and $14\%$ respectively compared to the baseline gray. Our results support the last hypothesis (colorfulness hypothesis) since all maps with colors increase accuracy over the gray baseline. %, as evidenced by the posthoc analysis: gray falls in a different posthoc group from. \begin{figure}[!tp] \centering \includegraphics[width=0.98\linewidth]{./figures1/EnsembleRanking.png}% \caption{Ensemble Ranking of Visualization Methods.} \label{fig:ranking} \end{figure} \subsection{Subjective Ratings and Comments} %\todo[inline, color=green!40]{ADD THE RATING FIGURES - GUOHAO has the numbers} %Figure~\ref{fig:} Participants' ratings and comments provide useful insights into how the usefulness of the colormaps was perceived. Participants' subjective rating of the usefulness of these colormaps, from high to low are: task1: coolwarm (5), extended-blackbody (4.96), blackbody (4.96), diverging (4.75), isoluminant-rainbow (4.4), and grayscale (3.7); task2: absolute (5.3), eigenmap (5.3), Boy's surface (4.8), and uniform uniform-gray (2). Grayscale in task 1 and uniform gray with no coloring was rated least useful for both tasks. %Comparing participants' ratings and the task performance with different colormaps shows that their preferences are indicators of good performance of visualization techniques. The ratings match relatively well with the ac- curacy data: the gray-scale and the baseline gray received lowest scores, while the blackbody and Boy’s surface maps received higher scores. %Overall, these ensemble tasks seems to be very challenging %and the highest mean ratings of these ensemble averages are about 5 for task type 1 or 6 for type 2 in a 7-scale. % The interviews revealed that those who liked the \textit{absolute} method found it the simplest to understand and easiest for following the tracts because of its symmetry; in addition, the less chaotic color changes helped them recognize the orientations better. Those who disliked the absolute method thought that tracts looked too similar to differentiate, show- ing the tradeoffs between similarity and resolution. Most participants were relatively neutral on the \textit{Boy's surface}, considering it similar to the \textit{eigenmap} method in terms of hue uses (spatial resolution) despite including more hues than that method. Participants commented that ``\textit{it (Boy's surface) was useful to have some different hues, but too many hues made the visualization less intuitive}'', while others stated that the ``\textit{right amount of hues of eigenmap provided enough discriminations between values without overloading one?s perception capability.}'' %It is not surprising that these comments match the accuracy observed in experiments: the rainbow perhaps was not bad because of the isoluminance but because of %many hues in the maps. \begin{figure*}[!th] \centering \includegraphics[width=0.99\linewidth]{./figures1/distributions.png}% \caption{Example Dataset Distribution and Their Colormaps: top: high-variance; middle: higher mean FA and narrow long-tail; bottom: low mean FA and narrow variance.} \label{fig:distribution} \end{figure*} \section{Discussion} This section discusses our results. Fig.~\ref{fig:ranking} shows our recommendations for choosing colormaps for the two ensemble tasks studied here. \begin{figure*}[htp] \centering \includegraphics[width=\linewidth]{./figures1/EigenBadCase7.png}% \caption{An example from the empirical study for which all participants got correct answers using absolute and Boy's surface but only half the participants got correct answers with Eigenmap. Red dots in the subfigures are sources. Eigenmap tends to show similar colors in cases in which the other two methods produce visually distinguishable ones. } \label{fig:eigenBadcase} \end{figure*} \subsection{Isoluminant Rainbow Does Not Decrease the Mean Accuracy, but Introduce Biases} Our first hypothesis is only partially supported. The most interesting result may be that the isoluminant rainbow does \textit{not} introduce greater error on average for task 1 (Fig.~\ref{fig:t1accuracy}). This efficiency result may agree with those in vision science because humans can average hues because humans can average hues{~\cite{maule2014getting}}. However, none of the vision science studies to our knowledge drills down to the empirical study results to examine whether or not participants would be biased towards higher or lower than ground truth. The fact that isoluminant rainbow introduces higher overshooting needs to be further studied, perhaps by explicitly controlling the variance in data for us to learn the colormap behaviors. Rainbow colors are known to be poor for univariate encoding due to the lack of uniformity and ordering and because they produce artificial boundaries in data. We could conclude from our study that ensemble color processing differs from univariate colormap representations. We do not recommend this isoluminant-rainbow map for ensemble average tasks. Instead, we propose to further explore \textit{how} and \textit{why} multihue works for limited capacity ensemble processing. This is mainly because the biases in isoluminant rainbow are consistent independent of the variances in data (Fig.{~\ref{fig:dataSpread}} and Fig.{~\ref{fig:biasTowardsHigher}}). The rainbow map certainly uses a set of semantically meaningful colors that would ease human understanding and our brain scientist collaborators particularly love rainbows; %In vision science, the study of subsampling of ensemble data is called \textit{limited capacity ensemble processing}. however, rainbow maps may still violate Trumbo's color design heuristics that \textit{``the basic information should be displayed in a clear and logical fashion so that it may be decoded with precision and without continual references to the key (labeled scheme)''} and \textit{``if small neighboring regions produce illusion of color over larger map areas, these illusions should not give misleading information''}~\cite{trumbo1981theory}. \subsection{Multihue Maps Improve Ensemble Accuracy in General} Our second hypothesis about multihue efficiency is supported. %In vision science, the access to low-variance local regions in colormap uses is called \textit{limited capacity ensemble processing.} %Zmultihues are unsuited to ensemble coding perhaps only in cases when the angular differences among hues approach parity ($180^\circ$){~\cite{webster2014perceiving}}. We ran a statistical analysis to examine whether or not hue or luminance affect error or task completion time. We found that hue had a significant main effect on time ($F(2, 1728) = 4.99$, $p = 0.0069$). The post-hoc analysis showed that colormaps with multihue led to statistically significantly longer task completion time than single-hue (gray) colormaps. %Using many hues may not help quantitative comparison because hues lack the natural ordering that luminance possesses. %Our brain data characteristics might only need an observer to see a small segments at a time. The multihue extended-blackbody and the coolwarm colormaps had the lowest absolute error, with slightly longer task completion time. This accuracy result of extended-blackbody agrees with 2D study results as well, though we did not observe significant differences. There may be at least two reasons for the benefits. First, one might think these two colormaps had the largest arc-length and thus yielded slightly better results than other maps. The other, perhaps primary reason for the benefits is that the multihue lets participants quickly determine the target-region first before formulating their answers, and this two-stage viewing could also explain why rainbows also take longer to execute. Visual inspection of colormaps applied to empirical data in three different FA distributions (Fig.~\ref{fig:distribution}) shows the FA variances when the mean is around the middle (top row), in the higher (middle row) or lower (bottom row) end. We may observe that the colormaps in the last three columns with many hues may help viewers quickly locate the target regions on the colormap into which the answers fall. \begin{comment} Our empirical study results strongly encourage so-called use-inspired design~\cite{wolfe2016use} in which real-world uses alter data distribution and task complexities. The artificial displays used in many vision and visualization science lab experiments tend to use reliable patterns that may have little relation to real-world data distributions. \end{comment} %Variation in hue tends to be represented as forming a circular perceptual continuum. %It is intuitive to think that the circularity of hue perception will interfere with the %ensemble coloring and as the angle differences approach $180^\circ$, %the competing interpretations of the mean color become equally likely and make the %mean extract more difficult. %Empirical study showed that this is not the case. \subsection{That Many Colormaps Work Well Also Shows the Power of Human Visual Systems in Judging Ensemble Averages} We did not observe differences in accuracy among colormaps when measuring the distance of participants' answers from the ground truth. This result suggests the power of visual ensembles for quantitative estimates. \begin{comment} Scientific data is often highly structured and may carry redundant structures resulting from sampling and filtering large simulations. The benefits of averaging depend on the extent to which the noise in individual datasets is correlated (less correlated, more benefit) and the number of individual measurements averaged (more measurements, more benefit). \end{comment} Balancing all considerations of efficiency, error, and correctness, and bias in these colormaps, we rank them in the order shown in Fig.~\ref{fig:ranking} task 1, where extended-blackbody, coolwarm, and blackbody seem to work well. Isoluminant-rainbow and diverging are worth further investigations. Gray is not recommended because of their higher biases. Though we cannot say whether the poor performance of grayscale was caused by its simultaneous contrast or its sole luminance channel, the result indeed is in agreement with the literature on 2D colorization. \subsection{Local Contrast and Resolution Together Might Be the Most Decisive Property for Ensemble Direction Tracing} Our results present an uncanny valley effect where the highest and lowest resolution maps improved outcome compared to the mid-resolution eigenmap orientation map. H4 is not supported. Overall, our results did not suggest that \textit{resolution} contributes to higher accuracy in 3D space, since both the Boy's surface and absolute methods reduced errors. The eigenmap had reasonable resolution, as does the Boy's surface colormap, but lowered accuracy. To understand \textit{when} Boy's surface and absolute succeeded and eigenamps failed. we inspected qualitatively by the best and worst examples of participant accuracy when using these colormaps, as shown in Fig.~\ref{fig:eigenBadcase}. We see that, while eigenmap provides regional coloring, the adjacent regions have relatively low contrast compared to other two approaches. These observations may suggest that local contrast is the most decisive property, since a combination of high contrast and spatial resolution, as in the Boy's surface, led to higher accuracy on ensemble tracing. Boy's surface generates %reasonably simple yet elegant colors that seem to strike the right balance in the spatial resolution and contrast for this spatial structure determination. Finally, the data sample varies so no dataset is seen twice by the same participants. For the eigenmap, this setting means that the colors for the same tracts in different datasets would change, while the same tube would always be given the same color with the other maps. We therefore recommend Boy's surface and absolute for coloring DMRI ensemble set, as shown in Fig{~\ref{fig:ranking}}. \subsection{Reuse of Our Results to Other Ensemble Representations} We sought to further our understanding of the ensemble data processing to generate concrete implications for visual analysis of brain DMRI tractography datasets. In general, both tasks suggest that high-contrast localized colormaps may have helped both ensemble average and tract discrimination. Reuse of our results in other domains would have to take into account domain specificities of data, task, and user. Several areas could benefit from our work, such as weather forecasting{~\cite{sanyal2010noodles}}, hurricane track prediction{~\cite{cox2013visualizing}}, and motion or movement trajectories{~\cite{chen2009visual}}{~\cite{andrienko2016leveraging}}, because direct trajectory depiction has been informative. The most suitable reuse would be when the datasets have relatively low variance, so that colormaps can be localized to a smaller regions on a colormap for scalar data visualizations. Similarly, the spherical orientation colormap for line field visualizations might also be domain-dependent. In our case, the tracts are following three major orientations. We also did not consider other tract shapes. Considering appropriate distance measures is needed for maximal performance. \subsection{Participants' Experiences} Participants in this study have different backgrounds, and an ideal condition might be to use only brain scientists, clinicians, or medical school students. One major reason for the background differences was that we had access to only a few brain scientists. We used as many as possible in the study because we wanted to collect their comments related to the brain science domain. Also, we followed Munzner's approach~\cite{munzner2009nested} of abstracting tasks into a level suitable for empirical study. In other words, these tasks could be performed by a trained participant. This may explain why we did not observe differences in task completion time and accuracy between students with and without medical backgrounds. Several user studies in flow visualization have used non-domain experts, suggesting that non-domain-expert is a viable option in empirical studies~\cite{liu20122d}. \subsection{Using Ensemble for Visualization Design} It is intuitive to think that hue, due to its categorical effect (e.g. yellow or red), would interfere with the ensemble coloring, thus making representing a multihue average difficult. However, this turns out not to be the case. In vision science, ensemble is believed to be used by the human visual system to address our severely limited visual working memory. We can quickly derive patterns that guide our attention towards the most useful information. Scientific data is often highly structured and may carry redundant structures. When there is redundancy, it is possible to sample and filter to produce optimal views. For example, a handful of past visualization work has shown that implicit or explicit representation of sets of objects as groups or ensembles can guide observers' attention to process only the most relevant incoming information (e.g., explicit depiction of a group of objects in clusters~\cite{peng2012mesh}, grouping interfaces to augment exploration workflows~\cite{li2011visbubbles}~\cite{ragan2016characterizing} or using spatial patterns to form texture pattern to guide observers' behavior~\cite{zhao2017bivariate}). We believe there will be an opportunity to create a compressed and efficient ensemble representation of information, such as ensemble overviews, to guide visual attention to the areas more relevant to the targets. \subsection{Limitations and Future Work} The aim of this paper was to investigate the effect of coloring in practice on two spatial ensemble visualization tasks, average and set orientation. Our study is only a first step towards understanding ensemble tasks in visualizations. Although this study can suggest \textit{what} colormap to choose for ensemble representation, we may need to build computational models or isolate factors (e.g., hue and luminance for task 1 and resolution and uniqueness for task 2) to explain \textit{how} these colormaps are used by our visual system. Effectiveness of these coloring approaches needs to be studied further when tasks are related to other discrimination and detection tasks, in which quantitative differences among data are to be reported. Our study would suggest further work. Since multihue colormaps in general improved ensemble average accuracy, one could run studies to systematically control the mean and variance of the ensemble datasets to model the ensemble performance. Viewers make a two-alternative forced-choice judgment about which visualization method contains the larger average value. Sensitivities are measured based on the differences between the values. A psychometric function fitted to the data reveals sensitivity to the discriminative threshold to measure accuracy. Using this method, we could answer questions about \textit{why} and \textit{when} multihue average will be effective and how variance influences the effectiveness and efficiency. \section{Conclusion} This study is the first (to our knowledge) to compare different color ensemble encodings for 3D DMRI tractography visualizations. Results from the study provide the following insights for choosing 3D tube coloring ensembles. \begin{itemize} \item The most interesting result was that the isoluminant-rainbow performed reasonably well, though it did lead to more reporting bias towards higher than ground truth values than other colormaps. \item Extended-blackbody, coolwarm, and blackbody are reasonably accurate for ensemble average in 3D. Our analysis showed that hue had much larger influence on error than luminance. \item Our study on the ensemble set orientation discrimination supports the proposition that having some colors is significantly better than no color at all. \item Colormaps with better orientation contrast (e.g., the Boy's surface and the absolute approach) are most desirable for ensemble set orientation discrimination tasks such as tract tracing. \end{itemize} \appendices \begin{figure*}[!t] \centering \includegraphics[width=0.98\linewidth]{./figures1/colorSpace.png}% \caption{Colormap Profile for Showing Scalars in Task 1 (AverageFA tasks) in the L*A*B* color space. L-planes from bottom to top are L=5, 20, 40, 60, 80, and 95.} \label{fig:colorSpace} \end{figure*} %\begin{comment} \begin{figure*}[!t] \centering \centering \includegraphics[width=0.98\linewidth]{./figures1/face.png}% \caption{Using Faces to Examine the Luminance Profile of Colormaps (from left to right): original image, gray, blackbody, diverging, isoluminant-rainbow, extended-blackbody, and coolwarm colormaps.} %\todo[inline, color=green!40]{INCOMPLETE. GUOHAO's COLOR MAPS HAVE NEGATIVE VALUES.} \label{fig:face} \end{figure*} %\end{comment} \section{The Univariate Colormaps in the L*A*B* Color Space} Fig.~\ref{fig:colorSpace} shows the scalar colormaps in the L*A*B* color space. The curve in each figure shows the trajectory of color maps and their three projects in the L*A*B* color space. All color interpolation is performed using linear interpolation in this space. %Our map permits variations in %This image samples 1000 points along the $L=0$ to $L=100$ plane. %\section{Colormap Profile Qualitative Inspection} %\section{The Blair-face Approach to Select Spiral Color Series} %Some colormaps has a range of hue and luminance. We used the Rogowitz-Kalvin{~\cite{rogowitz2001blair}} and Kindlmann-Reinhard-Creem approaches{~\cite{kindlmann2002face}} to help visually inspect colormaps to test their luminance profile. This method utilizes our sensitivity to luminance variations in human faces to select colormaps. Fig.{~\ref{fig:face}} shows samples of faces generated by these six colormaps with our online tool.The faces with isoluminance-rainbow and diverging colormaps are less recognizable than all others. The rainbow and coolwarm colormaps help distinguish different values: one can clearly see red (high) values around the nose and under the eyes. %where the colors %are mapped to the luminance of a human face. %We use an image with a broad range of gray scale as shown in Fig.~\cite{fig:face}. %\todo[inline, color=green!40]{ADD OBSERVATIONS MADE FROM FACES} \section{Coloring Tool Website} \begin{figure}[!tp] \centering \includegraphics[width=\linewidth]{./figures1/colorTool.png}% \caption{Exploratory Color Comparison Tool.} \label{fig:tool} \end{figure} Our own tool (Fig.{~\ref{fig:tool}}) is hosted at {http://wchiou1.github.io/colorTool/({Fig.~\ref{fig:tool}})}. During the evaluation process, we found that using a coloring tool to quickly provide side-by-side comparison made our discussion with the medical doctors very effective and efficient. The direct manipulation interface lets users directly drag and drop plain-text colormaps. It can display both 2D image and 3D geometry examples. % use section* for acknowledgment \ifCLASSOPTIONcompsoc % The Computer Society usually uses the plural form \section*{Acknowledgments} \else % regular IEEE prefers the singular form \section*{Acknowledgment} \fi The authors would like to thank Drs. Peter Kochunov, L. Elliot Hong, and Neda Jahanshad for their discussions on color uses in brain science, Dr. Bernice Rogowitz for a discussion on colormap uses in real-world applications, Dr. Jeremy Wolfe for his thorough review and comments on this manuscript, and the anonymous reviewers for their constructive comments. The authors also thank the participants at University of Maryland, Baltimore Country, University of Maryland Medical School, and Veterans Affairs Medical Center of Providence, RI for their time and effort. We thank Katrina Avery for her editorial support. This work was supported in part by NSF IIS-1302755, CNS-1531491, DBI-1260795, IIS-1018769, and DUE-0817106 and by NIST MSE-70NANB13H181. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of National Institute of Standards and Technology (NIST) or the National Science Foundation (NSF.) Jian Chen is the corresponding author. % Can use something like this to put references on a page % by themselves when using endfloat and the captionsoff option. \ifCLASSOPTIONcaptionsoff \newpage \fi %\todo[inline, color=green!40]{FIX THE REFs AND ADD DOI.} \bibliographystyle{IEEEtran} % argument is your BibTeX string definitions and bibliography database(s) \bibliography{main} % % <OR> manually copy in the resultant .bbl file % set second argument of \begin to the number of references % (used to reserve space for the reference number labels box) %\begin{thebibliography}{1} %\bibitem{IEEEhowto:kopka} %H.~Kopka and P.~W. Daly, \emph{A Guide to \LaTeX}, 3rd~ed.\hskip 1em plus % 0.5em minus 0.4em\relax Harlow, England: Addison-Wesley, 1999. %\end{thebibliography} % biography section % % If you have an EPS/PDF photo (graphicx package needed) extra braces are % needed around the contents of the optional argument to biography to prevent % the LaTeX parser from getting confused when it sees the complicated % \includegraphics command within an optional argument. (You could create % your own custom macro containing the \includegraphics command to make things % simpler here.) %\begin{IEEEbiography}[{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{mshell}}]{Michael Shell} % or if you just want to reserve a space for a photo: %\begin{IEEEbiography}{Michael Shell} %Biography text here. %\end{IEEEbiography} % if you will not have a photo at all: %\begin{IEEEbiographynophoto}{John Doe} %Biography text here. %\end{IEEEbiographynophoto} % insert where needed to balance the two columns on the last page with % biographies %\newpage %\todo[inline, color=green!40]{ADD AUTHOR BIO - PLEASE USE THE ONES FROM THE FMRI PAPER AND PLEASE COPY THE PHOTOS.} %\begin{comment} \begin{IEEEbiography} [{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{figures1/bio/Jian.png}}] {Jian Chen} received the PhD degree in Computer Science from Virginia Polytechnic Institute and State University (Virginia Tech). She did her post-doctoral work in the Department of Computer Science at Brown University. She is an Associate Professor in Computer Science and Electrical Engineering at The Ohio State University where she directs the Interactive Visual Computing Laboratory (IVCL). Her research interests include design and evaluation of visualization techniques and virtual reality. She is a member of the IEEE and the IEEE Computer Society. \end{IEEEbiography} \begin{IEEEbiography} [{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{figures1/bio/Zhang.png}}] {Guohao Zhang} is a PhD student in the Department of Computer Science and Electrical Engineering at University of Maryland, Baltimore County. He received his B.E. degree in Engineering Physics from Tsinghua University in 2012. His research interests include design and evaluation of visualization techniques and 3D visualizations. He is a student member of IEEE. \end{IEEEbiography} \begin{IEEEbiography} [{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{figures1/bio/Wesley.png}}] {Wesley Chiou} is an undergraduate student in the Department of Computer Science and Electrical Engineering at University of Maryland, Baltimore County. His research interest is human-computer interaction and visualization. \end{IEEEbiography} %\begin{comment} \begin{IEEEbiography} [{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{figures1/bio/David.png}}] {David H. Laidlaw} received the PhD degree in computer science from the California Institute of Technology, where he also did post-doctoral work in the Division of Biology. He is a professor in the Computer Science Department at Brown University. His research centers on applications of visualization, modeling, computer graphics, and computer science to other scientific disciplines. He is a fellow of IEEE. \end{IEEEbiography} %\end{comment} \begin{IEEEbiography} [{\includegraphics[width=1in,height=1.25in,clip,keepaspectratio]{figures1/bio/Auchus.png}}] {Alexander P. Auchus} Dr. Alexander P. Auchus holds degrees from Johns Hopkins University and from Washington University in St. Louis. He is an elected fellow of the American Neurological Association, the American Academy of Neurology, and the American Geriatrics Society. He has served on the faculty of Emory University, Case Western Reserve University, and University of Tennessee. His present position is Professor and McCarty Chair of Neurology at the University of Mississippi Medical Center. Dr. Auchus's research interests are in neuroimaging biomarkers for Alzheimer's disease and other dementias. \end{IEEEbiography} %\end{comment} %\begin{IEEEbiographynophoto}{Jane Doe} %\end{IEEEbiographynophoto} % You can push biographies down or up by placing % a \vfill before or after them. The appropriate % use of \vfill depends on what kind of text is % on the last page and whether or not the columns % are being equalized. %\vfill % Can be used to pull up biographies so that the bottom of the last one % is flush with the other column. %\enlargethispage{-5in} % that's all folks \end{document} }
\caption{Task 2: Mean Time and Accuracy. The color schemes connected by the \textcolor{orange}{orange} line are significantly different.}
\caption{(a) Ferromagnetic spin configuration aligned along the in-plane $y$-direction. The corresponding unit cells are represented by green-dashed rhombuses. (b) Coplanar 120$^\circ$ spin configuration. (c) Magnetization density isosurface at $n_\uparrow(\vecr)-n_\downarrow(\vecr) = 0.001\ e$/\AA$^3$. (d) Cross-sectional views of the magnetization density plotted along the $S1$ (top panel) and $S2$ (bottom panel) green-dashed lines introduced in (c). \red{(e) Uniaxial magnetocrystalline anisotropic energy $E_{\rm MAE}$ drawn in polar coordinates ($r,\theta,\phi$), where $r = E_{\rm MAE}(\theta,\phi)$. (f) Uniaxial magnetocrystalline anisotropic energy curves $E_{\rm MAE} = E_{\rm MAE} (\theta,\phi)$ evaluated as a function of $\theta$ at $\phi = 0$ (blue) and as a function of $\phi$ at $\theta = \pi/2$ (green), respectively. Inset illustrates the parametrization of spin orientation in terms of the azimuth angle $\theta$ measured from the out-of-plane axis and the horizontal azimuth angle $\phi$ measured on the basal plane. The red arrow represents spin orientation $\hat{\boldsymbol{n}}=(\sin\theta\cos\phi,\sin\theta\sin\phi,\cos\theta)$ in real space. } }
\caption{\red{(a) Electronic energy band structure and projected density of states (PDOS) of \textit{h}-InC. (Left panel) Spin-polarized electronic energy band structure. The Fermi level is set to zero. The energy bands are calculated without SOC along the high-symmetry $k$-points of the hexagonal BZ shown in \fig{fig:atomic}(b). The spin-up (majority spin) and spin-down (minority spin) bands are colored by red and blue, respectively. The mirror eigenbands with $m_z$ = +1 and $m_z=$ -1 are illustrated with solid and dashed lines, respectively. (Right panel) PDOS of \emph{h}-InC. The same energy scale is used between the band structure and PDOS. (b) Magnified view of the gray box in the left panel of (a). A band crossing is highlighted by a green circle. (c) Weyl nodal lines in energy-momentum space. Figure plots only the nodal lines within the energy range $|E| < 0.4\ $eV. The color scheme is used to represent the energy of the Weyl nodal lines. The node that is labeled by A (B) in (b) is a member of the type-II (type-I) Weyl nodal line in (c).}}
\caption{Contrasting Fermi surfaces of type-I and type-II nodal lines in \emph{h}-InC. (a) Fermi surfaces at $E$ = -0.238 eV, which intersect with a type-I nodal line. The orange hexagon illustrates the BZ. The Fermi surfaces for the spin-up and spin-down bands are colored by red and blue, respectively. The momenta of the nodal line are indicated by the dashed line. (b) Magnified black-boxed region from (a). Only the Fermi surfaces (contours) that intersect with the type-I nodal line are highlighted. (c) Magnified black-boxed regions from (b) (top panels) and the energy bands depicted along the green line from the corresponding top panel (bottom panels). (d) Fermi surfaces at $E$ = 0.346\eV, which intersect with the type-II nodal line. (e) Magnified black-boxed region from (d). Only the Fermi surfaces (contours) that intersect with the type-II nodal line are highlighted. (f) Magnified black-boxed regions from (e) (top panels) and the energy bands depicted along the green line from the corresponding top panel (bottom panels).}
\caption{Examples of the X-Rite ColorCheckers: a) ColorChecker\textregistered~Classic; b) ColorChecker \textregistered~Digital SG.}
\caption{Jackiw-Teitelboim black hole in global AdS$_2$ coordinates. The {\color{red} red curves} indicate the outer boundary of the 1+1-dimensional black hole spacetime, while the solid black curves indicate singularities where $\Phi=0$. The dashed diagonal lines show the location of the black hole horizon.}
\caption{Jackiw-Teitelboim black hole: Curves of constant Schwarzschild time outside the event horizon are shown in {\color{blue} blue}.}
\caption{Correlation-spectra corresponding to various Response Bits of {\em \color{blue} Correct} 5-4 DAPUF (for $N = 50$) \label{fig:correlation_plots}}
\caption{Correlation-spectra corresponding to various Response Bits of {\em \color{red} Faulty} 5-4 DAPUF (for $N = 50$) \label{fig:faulty_correlation_plots}}
\caption{ (Color online) (a) - (e) The maximal displacement $X_c$ (see Eq. \ref{eq-Xc}) as a function of $\beta$, $U$, $N$, $L$ and $\alpha$. (f) The area enclosed by the trajectory in phase space as a function of $\beta$. Parameters for open symbols \begin{scriptsize}\textcolor[rgb]{1,0,0}{$\bigcirc$}\end{scriptsize} are $L=21$, $N=10$, $U=\sqrt{2}$, $\alpha=1/(3\sqrt{5})$; \textcolor[rgb]{0,0,1}{$\square$} are $L=30$, $N=15$, $U=0.8$, $\alpha=0.1$ and \textcolor[rgb]{0,0.5,0}{$\bigtriangleup$} are $L=39$, $N=22$, $U=\sqrt{3}/2$, $\alpha=0.05$.}
\caption{\label{fig:The-potential-profile}The potential profile (left) showing the formation of inverse sheath (the wall is at $x=0$). The parameters used are $n_{e0}=10^{12}\,{\rm m}^{-3},T_{e}=1\,{\rm keV},J_{h\nu}=4.5\,\mu{\rm A/m^{2}},M=1.2,\sigma=1,\delta_{i}=1.1,\sigma_{{\rm ph}}=1$. The numbers in the figure indicates $\delta_{{\rm ph}}$. The panel on the right shows the normalized ion density as one moves from the wall. All the parameters are same with \textcolor{red}{$\delta_{{\rm ph}}=2.5$.}}
\caption{ The concentration-mass relation at $z=0$ (left) for the two modified families (halo 740 and halo 839 in purple squares and red diamonds respectively). Increasing the variance (arrows) in the initial conditions systematically yields more concentrated haloes. We compare these variations with correlations in the overall halo population (blue hexagon bins with median and 68 per cent confidence intervals in six bins of halo mass). Variations in the mass accretion history of two haloes generate changes in concentration that are comparable to the entire population-level scatter. These variations are consistent with the correlation between halo concentration and formation redshift (right), as both families move along the direction of the correlation. The jump in halo concentration between \ref{item:A} and~\ref{item:B} is tied to the reconfiguration of the merger tree shown in Fig.~\ref{fig:merger_reduction_740}. {\red{To verify this, we constructed an additional simulation (purple star), intermediate between the ends of the jump. We found that this version undergoes a three-way merger (i.e. it sits precisely on the transition between the two merger topologies) and as a result its concentration is also intermediate.}} }
\caption{Results of the Monte Carlo simulations with 15000 iterations per run (\textcolor{green}{competitive}, \textcolor{red}{collusive} and stakeless games).}
\caption{OTF vs ADM-RANS unsteady flow. The solid lines show the ADM-RANS results (left (\protect\tikz \protect\draw[color=red] (0,10)--(0.5,10);) and right (\protect\tikz \protect\draw[color=blue] (0,10)--(0.5,10);) ) and the dashed lines show the OTF results with a variable viscosity (left (\protect\tikz \protect\draw[style=dashed,color=red] (0,10)--(0.5,10);) and right (\protect\tikz \protect\draw[style=dashed,color=blue] (0,10)--(0.5,10);) ). The dotted lines show OTF results with $\nu=\SI{3}{\square\metre\per\second}$ (left (\protect\tikz \protect\draw[style=dotted,color=red] (0,10)--(0.5,10);) and right (\protect\tikz \protect\draw[style=dotted,color=blue] (0,10)--(0.5,10);) ). The shaded areas represent the range of values obtained if constant eddy viscosity values of $\SI{0.1}{\square\metre\per\second}\leq\nu\leq\SI{1}{\square\metre\per\second}$ are used in OTF.}
\caption{Optimised layout with cluster A turbines shown in circles (\protect\tikz \protect\draw[thick] circle (1.5pt);) and cluster B turbines shown in crosses (\protect\tikz \protect\draw[color=red,mark=x] plot(10cm,10cm);).}
\caption{Solar irradiance during the X9.3 flare of 2017 September 6 (with the pre-flare irradiance subtracted), observed by GOES (orange line) and LYRA channels 1, 2 and 4 (respectively the purple, green, and black lines for $E_1$, $E_2$ and $E_4$), as well as the Lyman-$\alpha$ residual irradiance $E'_1$ (red line) extracted from $E_1$. The LYRA data were rebinned to the cadence of 1 s. The time derivative of GOES 1--8~\AA\- data is also shown (blue line) as a proxy of the non-thermal flare emission. Different scales were used for the various time series for the sake of clarity.\label{fig:flare}}
\caption{Additional \aastex\symbols}
\caption{Figure taken from Ref. \cite{Quemeneur2014}: Protein lateral mobility in fluctuating membranes. Semi-logarithmic plot of the diffusion coefficients ($D_\mathrm{eff}$) as a function of the membrane tension $\Sigma$, for AQP0 (\textcolor{blue}{$\blacklozenge$}) and KvAP (\textcolor{red}{$\blacktriangle$}) labeled with streptavidin QDs. %Each point represents a median diffusion coefficient obtained from hundreds of individual trajectories for a GUV at a given tension; %the error bars correspond to standard error. KvAP data adjusted by Eq. \ref{effective D_lin} (solid line) yields a protein coupling coefficient $\Theta$ = 3.5$\times10^{-7}$ m considering $a = 5$ nm, $\kappa = 20$ k$_B$T and $D_0 \simeq$ 2.5 %$\upmu$ m$^2$/s. Simulations of the protein diffusion on a membrane subject to thermal fluctuations ($\blacksquare$) agree well with the experimental data and theory. Insets: sketches of membrane deformation near proteins.}
\caption{ Comparing peak performance on an Intel\textregistered{} Xeon\textregistered{} Scalable Platinum 8180 processor for TVM, ISAM-TVM and LIBXSMM when executing all ResNet-50 layers with the very small minibatch 28. }
\caption{(\textit{a}) Mean velocity profile, $U(y)$ and (\textit{b}) drag force, $D(y)$, scaled in outer units. Line styles are as defined in table~\ref{DNS_param}. \protect\reddashdotted, $D(y)$ for case H0F without accounting for the change in mean pressure gradient.}
\caption{\red{Success rate of Jacobian-based Data Augmentation attack against different defenses. RFN, SafetyNet, and Defense-GAN were evaluated on MNIST. Thermometer encoding was evaluated on CIFAR-10.}}
\caption{\red{Adversarial examples crafted for CIFAR-10 to fool a classifier fortified with thermometer encoding at different levels of perturbation.}}
\caption{\red{The success rate and average l2 norm of crafted adversarial examples by Liu et al. work against different defenses. RFN, SafetyNet, and Defense-GAN were evaluated on MNIST. Thermometer encoding was evaluated on CIFAR-10.}}
\caption{Visualizing the input and behavior of the fusion network. In the fusion indicator map, we compare the fused flow with two flow candidates. {\color{blue}{Blue}}: both candidates are similar ($<5$px), {\color{darkyellow}{yellow}}: the fused flow is similar to $\mv{w}_{t \rightarrow t+1}$, and {\color{cyan}{cyan}}: the fused flow is similar to $\widehat{\mv{w}}_{t \rightarrow t+1}$. {\color{red}{Red}}: the fused flow is different from all flow candidates (mostly occluded regions).}
\caption{Experimental SFs of various valence proton + doubly magic nuclei (\textcolor{lightblue}{$\times$}~\cite{Branford00}, \textcolor{lightblue}{$\bigcirc$}~\cite{Oliver69,Leighton68}, \textcolor{lightblue}{$\Diamond$}~\cite{Fortune75,Britton76}, \textcolor{green}{$\Box$}~\cite{Thoennessen03}, \textcolor{red}{$\bigcirc$} present work). Solid symbols represent the ground-state to ground-state channel. Open symbols represent the total strength of \orb{0}{d}{5} $\pi$-SPO. Shaded red circle is the sum of SFs of the \ch{23}{22} channel, which contains contributions from ground and excited states.}
\caption{Segmentation results on the validation set -- Top row shows Dice Similarity Coefficients and Hausdorff distance as a function of $\alpha$ in \eqref{eq:loss} for all three segmentation anatomies and both phases for $p=0.3$. Bottom row shows ground truth (left) vs. segmentation results (right) for $\alpha=0.05$ and $p=0.3$. From left to right, the cases are: NOR (correcly diagnosed), MINF (correctly classified even with artifact on top left), and ARV (misclassified as NOR). All show {\color{blue!90!black}LV in blue}, {\color{red!90!black}RV in red}, and {\color{green!60!black}Myo in green}. }
\caption{Runtimes on pre-compiled CryptoNets network with HE-SIMD packing for different batch sizes, for $N=2^{13}$ (\protect\blueline) and $N=2^{14}$ (\protect\redline).}
\caption{An example for a 3 way 2 shot scenario. Different colors indicate different entities, \textcolor{blue}{blue} for head entity, and \textcolor{red}{red} for tail entity.}
\caption{Examples from relation ``educated\_at''. Different colors indicate different entities,\textcolor{blue}{blue} for head entity, and \textcolor{red}{red} for tail entity.}
\caption{{\color{red}Classification accuracy on 10 probing tasks. ``\#L'' denotes the layer number. ``A1'', ``A2'' and ``A3'' represent aggregation node 1, 2, 3, respectively. `SeLen' is to predict the length of sentences in terms of number of words. `WC' tests whether it is possible to recover information about the original words given its sentence embedding. `TrDep' checks whether an encoder infers the hierarchical structure of sentences. In `ToCo' task, sentences should be classified in terms of the sequence of top constituents immediately below the sentence node. `Bshif' tests whether two consecutive tokens within the sentence have been inverted. `Tense' asks for the tense of the main-clause verb. `SubNm' focuses on the number of the subject of the main clause. `ObjNm' tests for the number of the direct object of the main clause. In `SOMO', some sentences are modified by replacing a random noun or verb with another noun or verb and the classifier should tell whether a sentence has been modified. `CoIn' benchmark contrains sentences made of two coordinate clauses. Half of the sentences are inverted the order of the clauses and the task is to tell whether a sentence is intact or modified.}}
\caption{ \label{fig:window_size} Distribution of the local scopes learned by each attentive layer. The upper figures illustrate the distribution of the predicted pairs of central position (Y-axis) and its correspond window size (X-axis) in each layer, the samples are randomly selected from the development set. The lower figures show the distribution of the window size in each layer. Blue color represents the \textcolor{blue}{layer-specific parametric} approach, while the \textcolor{red}{query-specific parametric} method is indicated in red. }
\caption{Fall semester daily average minutes ceiling light usage compared to weekday \& weekend average baselines. Vertical black dashed lines indicate a weekend period}
\caption{Tensor-network schematics of the formalism. (a) The single-qubit measurement $\boldsymbol{M}=\{M^{(a)}\}_a$ is a three-index tensor represented by a yellow circle with three emerging lines. Vertical indices act on the physical degrees of freedom while the outcoming one labels the measurement outcome $a$. (b) The $N$-qubit measurement $\boldsymbol{M}=\big\{M^{(a_1)}\otimes M^{(a_2)}\otimes \hdots M^{(a_N)}\big\}_{a_1, \hdots a_N}$ corresponds to the same local measurement on each qubit. (c) The components $T_{\boldsymbol{a},\boldsymbol{a}'} = \text{Tr}\left[ M^{(\boldsymbol{a})} M^{(\boldsymbol{a}')} \right]$ of the overlap matrix $\boldsymbol{T}$ correspond to the physical-index contraction of the measurement operators. (d) Consequently, when $\boldsymbol{T}$ is invertible, its inverse $\boldsymbol{T}^{-1}$ factorizes into the tensor product of single-qubit matrices $\boldsymbol{T}_{\text{1-qubit}}^{-1}$ (green squares). (e) We represent the outcome distribution $\boldsymbol{P}$ as an $N$-index tensor where the indices encode the measurement outcome of each qubit. (f) A generic density matrix $\varrho$ is given by a contraction of $\boldsymbol{P}$, $\boldsymbol{T}^{-1}$, and $\boldsymbol{M}$ over the outcome indices: $\varrho = \sum_{\boldsymbol{a},\boldsymbol{a}'} P(\boldsymbol{a}) \,T^{-1}_{\boldsymbol{a},\boldsymbol{a}'}\,M^{(\boldsymbol{a}')}= \mathbb{E}_{\boldsymbol{a}\sim\boldsymbol{P}}\left( \sum_{\boldsymbol{a}'} T^{-1}_{\boldsymbol{a},\boldsymbol{a}'}\,M^{(\boldsymbol{a}')} \right)$. Expressed this way, all non-local correlations of $\varrho$ are encoded explicitly into $\boldsymbol{P}$, as both $\boldsymbol{T}$ and $\boldsymbol{M}$ are single-qubit factorable. Our state Ansatz parameterizes the outcome distribution with a neural-network generative model, i.e. $\boldsymbol{P}= \boldsymbol{P}_{\text{model}}$. \label{fig:TNrep}}
\caption{Classification results. In \textcolor{green}{green} (\textcolor{blue}{blue}) best respectively worst F-measure and true positive results. Best balanced results highlighted in \textcolor{orange}{orange}. Overall best precision, recall and F-measure results are set in bold face.}
\caption{Unknown words ranked by SVM probability. Classification obtained with \textit{form, lex, theme} features. In parentheses: if validated (\textcolor{red}{$+$}) or not.}
\caption{\label{fig:DCNtemporalgraph} Temporal digraph of the DCN $\mathcal{C} := \{(1,4,\tau_1), (5,4,\tau_2), (2,5,\tau_3), (4,3,\tau_4)\}$ with $\tau_1 < \tau_2 < \tau_3 < \tau_4$. Note that there is a temporally coherent path from $\mathcal{C}$-vertices 1 to 3 (indicated with {\bf bold} versus {\color{gray}gray} arrows), but not from 2 to 3. Spatial (resp. temporal) arcs are vertical (resp. horizontal); temporal fibers are vertices along horizontal paths.}
\caption{\label{fig:SimpleMarkov} Entries of $P_{(\mathcal{C},a',1)}^{(\beta,\varepsilon)}$ not in $\{0,1\}$ for $\mathcal{C} := \{(1,4,1), (5,4,2), (2,5,3), (4,3,4)\}$ and $a' = \{0,5\}$ are indicated along with solid arcs ({\color{gray} gray} for spatial arcs; black for temporal arcs); unit entries correspond to dashed arcs.}
\caption{Performance for PED noise, {\bf all SNRs} averaged; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Performance for CAFE noise, {\bf all SNRs} averaged; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Performance for {\bf unseen} BUS noise, {\bf all SNRs} averaged; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Performance for PED noise at {\bf $\text{SNR}\mathbf{=-5}\,\text{dB}$}; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Performance for CAFE noise at {\bf $\text{SNR}\mathbf{=-5}\,\text{dB}$}; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Performance for {\bf unseen} BUS noise at {\bf $\text{SNR}\mathbf{=-5}\,\text{dB}$}; $\Delta$SNR and SSDR are measured in dB. {\bf Two best are \colorbox{black!15}{greyshaded}}.}
\caption{Experimental (bullets) and numerical (full lines) velocity plots at location A1 (see Fig.\ref{fig:2}) for the primary (\textit{a}) and wall-normal velocity (\textit{b}), for increasing values of $De^*$ (\textcolor{black}{$\bullet$} Newtonian fluid, \textcolor{green1}{$\bullet$} 0.43, \textcolor{orange1}{$\bullet$} 0.71, \textcolor{red}{$\bullet$} 0.86 and \textcolor{black}{\textbf{-}} 0, \textcolor{purple}{\textbf{-}} 0.11, \textcolor{blue2}{\textbf{-}} 0.22, \textcolor{green3}{\textbf{-}} 0.44, \textcolor{orange2}{\textbf{-}} 0.67, \textcolor{red}{\textbf{-}} 0.89). For easier comparison, only select $De^*$ values are shown in \textit{(b)}. The zero $y$ location is taken at the inner edge of the bend, with the geometry being that described in section~\ref{sec:expgeom}: $R_i/W = 0.36$. The peak streamwise velocity shifts with $De$, as is more visible in the inset of \textit{(a)}. The scale bar in the top right corner indicates the magnitude of the error on the experimental data. (\textit{c}) Scaling for the maximum wall-normal velocity with $De^*$.}
\caption{ How many zebras are there? \textcolor{green}{GT}: 5}
\caption{ How many people are pictured? \textcolor{green}{GT}: 4 }
\caption{How many dogs are there? \textcolor{green}{GT}: 2 }
\caption{ How many zebras are standing in this image? \textcolor{green}{GT}: 1 }
\caption{ How many people are not wearing red shirts? \textcolor{green}{GT}: 4}
\caption{How many people wear hat? \textcolor{green}{GT}: 1 }
\caption{\label{fig:consensus}\footnotesize \hyphenpenalty=10000 \exhyphenpenalty=10000 {\bf Segmentation consensus}. Results of interactive segmentation (columns 2 to 6) are shown for four small tiles (column 1) of much larger images. Although users create different segmentations for the same image, they nevertheless exhibit some agreement most prominent in clear, unambiguous parts of the image. Using our median approach to combine users' segmentations into a single one eliminates outliers and spurious annotations, \eg\dangling scribbles and speckles, as shown in our results on the last column, 8. Our solution is on par or better (third row) than results obtained with the Staple method\cite{warfield2004}, shown on column 7, and orders of magnitude faster (\eg\1.06s against 237.35s for sixteen 800x800 images) when compared to the expectation maximization approach adopted by Staple. Users' segmentations are from our{\em Collaborative Segmentation} web application. $h = 5\ \text{or}\ 10$.}
\caption{\label{fig:distance}\footnotesize{\bf Illustration of planar shape distances}. Arrows in the left figure represent the shortest distances $d_{\ct{}}(x)$ and $d_{\Psi}(y)$ from points $x\in\Psi$ and $y\in\ct{}$, respectively, to closed contours $\ct{}$ and $\Psi$. The figure on the right illustrates a case where the horizontal black curve $\ct{}$ is equally distant to three different other curves, $d(\ct{},\ct{1}) = d(\ct{},\ct{2}) = d(\ct{},\ct{3}) = 63$. The symmetric distance $\dg$ captures our two-way expectations: $\dg(\ct{},\Gamma_1) = 145 > \dg(\ct{},\ct{2}) = 129 > \dg(\ct{},\ct{3})= 116$. Curves $\ct{i}$ only differ at the bulgy part. Slanted arrows on the right figure are directed to the farthest points on $\ct{i}$ from $\ct{}$. To compute $\dg$, curves were 8-connected discretized in 429x295 images.}{\color{lightgray}\hrule}
\caption{Performance comparisons with the state-of-the-art predictors. \textcolor{blue}{Blue numbers} indicate the best performances achieved in the first-stage regression; \textcolor{red}{Red numbers} indicate the best performances after the second-stage regression.}
\caption{\small An illustrative example of soft matching in NPRF. The target document FBIS3-23332, judged relevant, is ranked 122$^{nd}$ by BM25 for query 341 on Robust04, and is promoted to the 5$^{th}$ by NPRF$_{ds}$-DRMM. The NPRF mechanism increases the chances of soft-matching query-related terms that appear in the top-ranked documents (terms in \textcolor{blue}{blue}), but are missing in both the query and the target document. Subsequently, the matching signals with the query terms (in {\bf bold}) and the query-related terms (in \textcolor{red}{red}) in the target document are enhanced. }
\caption{\label{fig:results}{\bf Highest scores for each dataset and algorithm.} The first row is the adjusted RAND index and the second row the adjusted mutual information. The highest score of the row is in {\bf \textcolor{darkturquoise}{green}} while the second highest is in {\bf \textcolor{darkorchid}{orange}}. The standard error over 10 runs is given in parentheses for DBSCAN++ with uniform initialization. Both other algorithms are deterministic. Each algorithm was tuned across a range of $\epsilon$ with minPts $= 10$. For both DBSCAN++ algorithms, we used $p$ values of $0.1, 0.2,$ or $0.3$. We see that DBSCAN++ uniform performs better than DBSCAN on 17 out of 22 metrics, while DBSCAN++ $K$-center performs better than DBSCAN on 21 out of 22 metrics.}
\caption{ (Color online.) Comparison between the usual DSM ({\color{Fuchsia}{\scriptsize $\blacksquare$}}), ReDSM ({\color{ForestGreen}{\large$\bullet$}}), and MUB QST ({\color{Orange} $\bm{\times$}}
\caption{ (Color online.) Comparison between the usual DSM ({\color{Fuchsia}{\scriptsize $\blacksquare$}}), ReDSM [including SSB ReDSM ({\color{ForestGreen}{\large$\bullet$}}) and BBB ReDSM ({\color{SkyBlue}{\large$\circ$}})], and MUB QST ({\color{Orange} $\bm{\times$}}
\caption{ (Color online.) Comparison between the usual DSM ({\color{Fuchsia}{\scriptsize $\blacksquare$}}), ReDSM [including SSB ReDSM ({\color{ForestGreen}{\large$\bullet$}}) and BBB ReDSM ({\color{SkyBlue}{\large$\circ$}})], and MUB QST ({\color{Orange} $\bm{\times$}}
\caption{Simulation results for setting 2, continuous $X$: relative root mean squared error (RRMSE) as a function of the outcome prevalence for the three estimands and the three sample sizes ($\bullet$ exact estimators, \textcolor{blue}{$\Box$} approximate estimators, \textcolor{red}{$\ast$} imputation, \textcolor{green}{$\bigtriangleup$} weighting).\label{fig:estcont}}
\caption{Simulation results for setting 2, binary $X$: relative root mean squared error (RRMSE) as a function of the outcome prevalence for the three estimands and the three sample sizes ($\bullet$ exact estimators, \textcolor{blue}{$\Box$} approximate estimators, \textcolor{red}{$\ast$} imputation, \textcolor{green}{$\bigtriangleup$} weighting).\label{fig:estbin}}
\caption{Simulation results for setting 2, continuous $X$, $n=1000$: empirical cumulative density function of estimated standard errors (--- exact estimators, \textcolor{blue}{---} approximate estimators, \textcolor{red}{---} imputation, \textcolor{green}{---} weighting).\label{fig:se}}
\caption{F1 scores (\%) obtained by our systems against the state of the art on the English WSD tasks of the evaluation campaigns SensEval 2 (SE2), SensEval 3 (SE3), SemEval 2007 (SE07) task 7 and 17, SemEval 2013 (SE13) task 12, SemEval 2015 (SE15) task 13 and the corpus composed of the concatenation of all previous ones (ALL) except SE07 task 7. Results in \tbf{bold} are the best results from using the sense vocabulary reduction or not. Results in \textcolor{red}{red} are to our knowledge the best results obtained on the task. Our results are the mean scores of 20 individual systems, with the standard deviation given in parenthesis. Results prefixed by a star (*) was obtained on the development corpus used during the training.}
\caption{F1 scores (\%) obtained by our system with an ensemble of 20 models trained separately. Results in \tbf{bold} are the best results from all our systems. Results in \textcolor{red}{red} are to our knowledge the best results obtained on the task.}
\caption{Three-dimensional spherically averaged energy spectrum of the initial condition: \texttt{kf02-a01} (\opensquare), \texttt{kf04-a01} (\redline), \texttt{kf08-a01} (\greycross), \texttt{kf16-a01} (\orangeline), \texttt{kf32-a01} (\blackline \opencircle \blackline), \texttt{kf04-a32} (\bluestar) \texttt{kf08-a16} (\opengreytriangleup) }
\caption{Time evolution of box-averaged kinetic energy (a) and energy dissipation rate (b) for $\romanrossbyeps\approx0.31$ (weak rotation). Lines corresponding to same $\wndpp$ are grouped by color: $\wndpp=2$ (\filledblacksquare), $\wndpp=4$ (\filledbluesquare), $\wndpp=8$ (\filledredsquare), $\wndpp=16$ (\filledorangesquare), Lines corresponding to the same $A_r$ are grouped by line types: $A_r=1$ (\blackline\opencircle\blackline), $A_r=8$ (\blackdashedtwoline), cf. \cref{tb:perfosimul}.}
\caption{ Time evolution of box-averaged kinetic energy (a) and energy dissipation rate (b) for $\romanrossbyeps\approx0.06$ (strong rotation). Lines corresponding to same $\wndpp$ are grouped by color: $\wndpp=2$ (\filledblacksquare), $\wndpp=4$ (\filledbluesquare), $\wndpp=8$ (\filledredsquare), $\wndpp=16$ (\filledorangesquare), $\wndpp=32$ (\filledgreysquare). Lines corresponding to the same $A_r$ are grouped by line types: $A_r=1$ (\blackline\opencircle\blackline), $A_r=2$ (\blackline), $A_r=4$ (\blackdashedline), $A_r=8$ (\blackdashedtwoline), $A_r=16$ (\blackdashedthreeline), $A_r=32$ (\blackline\opensquare\blackline), cf. \cref{tb:perfosimul}. }
\caption{Spectral energy flux for $\romanrossbyeps\approx0.06$ and cases with $\wndpar=32$ (a) and $\wndpp=16$ (b). In (a), $\wndpp=8$ (\redline), $\wndpp=16$ (\orangeline) and $\wndpp=32$ (\greyline). In (b), $\wndpar=16,32$ and $64$ (\orangeline). Arrow denotes the direction of increase.}
\caption{Three-dimensional spherically averaged energy spectrum for $\wndpar=32$ (a) and $\wndpp=16$ (b) with $\romanrossbyeps\approx0.06$. Line styles are the same as in \cref{fig:specfx}, apart from the reference energy spectrum of \cref{fig:iso_energyspec} with $\wndpp=\wndpar=32$ (\greydashedtwoline).}
\caption{ One-dimensional energy spectra for $\romanrossbyeps\approx0.06$ and $\wndpp=\wndpar=32$ (case \texttt{kf32-a01}) along directions $\wnpp$ (a) and $\wnpar$ (b). Lines represent the time evolution of the energy spectrum: $t=0$ (\blueline), and $t=10, 20$ and $30\,\tf$ (\blackline). A reference line for the scaling laws that best agrees with the presented data is also shown (\blackdashedline). }
\caption{MT from English into Portuguese varieties. Example of mixed translations generated by Google Translate (as of 20th July, 2018) and translations generated by our variety-specific models. For the underlined English terms both their \textcolor{olive}{Brazilian} and \textcolor{blue}{European} translation variants are shown.}
\caption{ English to Portuguese translation generated by Google Translate (as of 20th July, 2018) and translations into Brazilian and European Portuguese generated by \replace{one of our semi-supervised multilingual model ({\tt M-C2})}{our semi-supervised multilingual ({\tt M-C2} and {\tt M-C3\_L}) and supervised {\tt Spec} models}. For the underlined English terms both their \textcolor{olive}{Brazilian} and \textcolor{blue}{European} translation variants are shown. }
\caption{Graphic representation for both models, {\color{blue} with $q=P_{cc}$ and $p=P_{ff}$}}
\caption{{\color{blue}Graphic showing the region where the PLR is equal after $n$ iterations}}
\caption{{\it Transfer-Learning; (left) from an initial NMT model to a new language pair, model is applied after inserting the new vocabulary entries, for instance, the initial model \textcolor{gray}{$L_{n-1}$} parameters are transfered to \textcolor{black}{$L_n$} with the updated embedding space (i.e., keeping $V_p^1$, $V_p^2$ as overlapping entries, while replacing the non-overlapping $V_p^i$ with $V_c^j$ new language vocabularies), and (right) from an initial model \textcolor{gray}{$L_{n-1}$} to \textcolor{black}{$L_{n}$}, but incorporating both the previous and new language pair data and vocabulary entries.}}
\caption{Construction of the coding network $G^*$ from the I/O graph $G$. Figure (a) shows the I/O graph $G$ for an array $A$ consisting of $3$-bit strings $d_1, \cdots, d_8$ with $w=3$ and $B=2$. Figure (b) shows the coding network $G^*$ derived from $G$. All \textcolor{figurePink}{pink} edges have the capacity of $w$ bits and all \textcolor{figureBlue}{blue} edges have the capacity of $b$ bits. Capacity of other edges are specified by their labels. In this example we assume that $\pi(7)=1$ and $\pi(5)=2$. Therefore, the output block containing $C[\pi(7)]$ and $C[\pi(5)]$ is the first output block which is the node $9$, and we added an edge with the capacity $w$ from the node $9$ to the sinks $t_7$ and $t_5$. }
\caption{\label{fig:fpga_scheme_lock_ctrl} Lock-control links to PIDs and Ramp instruments. \textcolor{darkred}{Red lines} are signal data buses and \textcolor{turquoise}{turquoise lines} represents sets of buses for parameters configuration. All the Lock-in outputs and the ADC inputs can be used in the PIDs for signal processing and in the Lock-control for event identification and trigger.}
\caption{\label{fig:abs-exp-data} Absorption spectroscopy of Rubidium. The transmitted intensity across the Rubidium cell is plotted for a complete Ramp scan measurement (\blue line) and for a lock-start event (\orange line). The curve dips corresponds to ${}^{87}\text{Rb}$ and ${}^{85}\text{Rb}$ hyperfine absorption lines. }
\caption{\label{fig:Sat_abs_F3} Relevant signals for a saturated absorption spectroscopy open-loop scan (\blue) and a triggered closed-loop stabilization scheme to an peak maximum (\orange). \texttt{in2} is the transmitted signal sensed by the photodiode. \texttt{in1} is \texttt{in2} DC-decoupled and amplified (see figure~\ref{fig:abs-exp}), and is used for lock-in demodulation. \texttt{F1o} and \texttt{F3o} is the demodulated signals at 1f and 3f respectively. \texttt{ctrl\_A} is the control signal, running a Ramp scan. The demodulated signals show the first and third derivative-like behavior, with zero-crossing points on \texttt{in2} peaks maximum. \texttt{F3o} was used for \texttt{error} signal and the \red line shows the level threshold for lock-start trigger. The left $V_\texttt{signal}$ axis represent the voltage scale of this signals in the photodiode output. }
\caption{\label{fig:PDH_allan_dev} Allan deviation~\eqref{eq:allan-dev} of the fractional frequency of the laser derived from \texttt{error} signal (\blue), and from control signals (\orange and \green). The first one characterize the performance of the closed-loop stabilization system on different time ranges. The other two are plotted for qualitative comparison: the control signals Allan deviations for time ranges $\tau >> \tau_i$ represents the corrections made to the system to keep it locked, and can be used as a rough estimation of the open-loop behavior the system would have had without stabilization. Dashed lines mark the time constans $\tau_i$ for each PID integrator. At 10~seconds the stabilization scheme achieves an improvement of 3 orders of magnitude. }
\caption{\label{fig:PDH_relock} Relocking feature of Control-lock instrument. In this example the relock system was configured to start when \texttt{in2} signal falls below $3000\,\text{int}\,\cong\,366\,\text{mV}$ (\gray~line). The triangular sweep increase the scan amplitud on each semi-period until it reachs the lock condition again. This feature is described in subsection~\ref{subsec:lock-control_intrument}. }
\caption{\label{fig:atari-all} Learning curves of the 49 Atari games. The x-axis is training steps, and the y-axis is the mean return of the most recent 1,000 episodes up to the time step. Blue lines indicate {\color{blue} QUOTA}, green lines indicate {\color{green} QR-DQN}, and red lines indicate {\color{red} QR-DQN-Alt}. All the scores are averaged over 3 independent runs, and standard errors are plotted as shadow.}
\caption{Evaluation curves of Roboschool tasks. The x-axis is training steps, and the y-axis is score. Blue lines indicate {\color{blue} QUOTA}, green lines indicate {\color{green} DDPG}, and red lines indicate {\color{red} QR-DDPG}. The curves are averaged over 5 independent runs, and standard errors are plotted as shadow.}
\caption{Collaborative filtering based positive pair mining. Given an query image $p$ (\textcolor{blue}{blue}) of real data, we first compute the $k$-reciprocal nearest neighbors $R_k(p)$ of $p$ (\textcolor{green}{green}). Then, the collaborator set (\textcolor{blue}{blue}) of $p$ and each candidate $q$ in $R_k(p)$ is mined in step (b). The collaborative filtering similarity of $p$ and each candidate $q$ in $R_k(p)$ is calculated by Eq.~\ref{eq:CF} in step (c). Finally, image pair with the highest re-calculated similarity is selected as the positive pair (\textcolor{green}{green}) in step (d).}
\caption{\label{baseline}Comparison of our MSC features with both the hand-crafted features and deep convolutional features within the DCF framework on OTB-2013. Note that $^{*}$ indicates the GPU speed, otherwise the CPU speed. The {\color{red} first}, {\color{green} second} and {\color{blue} third} best features are shown in color.}
\caption{\label{overallTable-MSCDCF}DPRs (\%) and speed ($^{*}$ indicates the GPU speed, otherwise the CPU speed) obtained by our MSC-DCF and MSC-CCO trackers as well as the state-of-the-art real-time trackers on OTB datasets. The results of top 8 performing trackers are given. The {\color{red} first}, {\color{green} second} and {\color{blue} third} best trackers are shown in color.}
\caption{\label{overallTable-MSC-CCO}OSRs (\%) and speed ($^{*}$ indicates the GPU speed, otherwise the CPU speed) obtained by MSC-CCO and MSC-DCF as well as the state-of-the-art deep feature-based CF trackers on OTB-2013, OTB-2015 and OTB-50. The {\color{red} first}, {\color{green} second} and {\color{blue} third} best trackers are highlighted in color.}
\caption{\textcolor{red}{Similarity estimation accuracy\protect\\ comparison between real value and\protect\\theoretical value on Last.fm.}}
\caption{\textcolor{red}{Similarity estimation accuracy\protect\\ comparison between real value and\protect\\theoretical value on MovieLens 20M.}}
\caption[]{Diphoton signals published by CMS \cite{CMSPASHIG-13-016} ({\definecolor{tmpclr}{rgb}{1.000,0.440,0.030}{\color{tmpclr}$\bullet$}}) and ATLAS \cite{ARXIV13053315} ({\color{blue}$\bullet$}), four-lepton signals by CMS Collaboration \cite{PRD89p092007} ({\color{green}$\star$}) and ATLAS \cite{ARXIV13053315} ({\definecolor{tmpclr}{rgb}{0.000,1.000,1.000}{\color{tmpclr}$\bullet$}}), invariant-mass distributions for $\tau\tau$ in $e^{+}e^{-}\to\tau\tau (\gamma )$ ({\definecolor{tmpclr}{rgb}{0.500,0.000,1.000}{\color{tmpclr}$\diamond$}}) and $\mu\mu$ in $e^{+}e^{-}\to\mu\mu (\gamma )$ ({$\bullet$}) by L3 \cite{PLB479p101}. }
\caption{Free Body Diagram. Illustrating all forces that impede the motion of a robot. The cuboid shown represents the robot with all forces acting around its center of mass. \textcolor{deepskyblue}{$W_R$}$ (= m_R g)$ represents the weight of the robot. We decomposed the weight components into a parallel component \textcolor{deepskyblue}{$W_{R_{\parallel}}$}$ (= m_R g sin\theta)$ acting along the terrain and perpendicular component \textcolor{deepskyblue}{$W_{R_{\perp}}$}$ (= m_R g cos\theta)$ acting against the \textcolor{orange}{$Normal (N)$} which represents the normal force. \textcolor{red}{$Friction$} represents the surface friction offered by the ground and \textcolor{darkblue}{$F_{air}$}$=cv^2$ represents the aerodynamic drag force.}
\caption{Energy distribution from battery pack. Illustrating how the energy from the battery is distributed across various \textcolor{red}{\textbf{Maneuvering}} and \textcolor{green}{\textbf{Ancillary}} components. $W_j$ represent the work done by various components. }
\caption{\label{fig:modes}Spatial distribution of the average displacements $e_i^2$ on the unstable modes for three stationary points sampled in the ternary mixture (a) well above ($T=0.45$), (b) around ($T=0.35$) and (c) at the mode-coupling temperature ($T=0.29$). Only particles in a vertical slab of thickness $2\sigma$ are shown. The color coding interpolates between yellow (\textcolor{snapyellow}{\large\textbullet}) for particles with $e_i^2=0$ and red (\textcolor{snapred}{\large\textbullet}) for those with $e_i^2\ge 0.04$.}
\caption{a) Global spatial distribution of Lagrangian drifters \red{with a spatial resolution of one degree both latitudinally and longitudinally}. The concentration of the Lagrangian drifter is partially due to the flow of the ocean current. For display convenience, the number $N$ has been taken its logarithm. A square (dashed line) indicates a area $x\in[150,250]$ and $y\in [-40,40]$ to exclude the continent boundary influence of calculating fractal dimension, see Fig.\,\ref{fig:FD}. b) The time evolution of the number of Lagrangian drifters on the time span 1st Jan. $2000$ to 1st Jan. $2012$. The number $N$ approximates to a constant 1,200 since 1st Jan. $2006$. The inset shows the pdf of the life of Lagrangian drifters, in which the $\sigma\simeq 360$ is the standard deviation. An exponential law is observed with a scaling exponent $0.40\pm0.02$ in the range $0.5\le T/\sigma\le 5$. }
\caption{a) The experimental second-order structure-function $S_2(\tau)$ in the range $2<\tau<100\,$Dtay (corresponding to $0.01<f<0.5\,$Day$^{-1}$) with a scaling exponent $\zeta_{\tilde{\theta}}^S(2)=1.36\pm0.02$, which is slightly larger than the prediction by the Fourier spectrum, i.e. $2.32-1=1.32$. \red{To clarify the scaling behavior, the first few points are ignored here}. The inset shows the compensated curve to emphasize the observed power-law behavior. b) The relative cumulative function $\mathcal{R}_{\theta}({f_M},\tau)$ with $f_M=f_Y=1/T_Y$. Note that the second-order SF is strongly influenced by the scales larger than annual cycle. The power-law range predicted by the Fourier analysis is indicated by a vertical line. }
\caption{Comparisons of the focal length estimated from three schemes on synthetic data: randomly taken images, calibration wizard and wizard using autocorrelation matrix. $f = 800, (u,v) = (320,240), k_1 = 0.01, k_2 = 0.1$. Initial calibration was done with $3$ random images. Left: Mean values of the estimated focal length, where the {\color{red} red} dashed line represents the ground truth $f=800$. Right: Standard deviations of the estimated focal length. Wizard images provide significantly more accurate and precise calibration results than random ones.}
\caption{Pose estimation test. Left: checkerboard image where $4$ {\color{green}green} corner points are used for pose estimation, and the remaining $50$ {\color{red}red} points for reconstruction. Right: $50$ ground-truth points in black and residuals between them and the reconstructed corner points in {\color{red}red} (enlarged 50 times for visualization).}
\caption{Comparisons of the intrinsic parameters estimated from three schemes on synthetic data: the paths generated from random poses, generated from the poses proposed by calibration wizard without and with autocorrelation matrix. $f = 800, (u,v) = (320,240), k_1 = \mathbf{0.5}, k_2 = \mathbf{1}$. {\color{red} Red} dashed lines represent the ground truth values. Wizard images achieve superior performance over random images on all intrinsic parameters.}
\caption{\blue{ Comparison of state evolution and simulation results under SNR = $10$ dB in (a) and $30$ dB in (b). $N = 256$. The results are averaged by $1000$ realizations.}}
\caption{\blue{NMSE performance of various algorithms under $\text{SNR}\in[-10,30]$ dB and pilot numbers $M\in[0.05N,N]$. A $21 \times 20$ grid is constructed for each algorithm with the performance averaged by $100$ independent trials at each grid point.} }
\caption{\blue{NMSE of various algorithms versus the number of pilot sequences $M$ under SCM, with $N = 256$ and $\text{SNR} = 10$ dB. The algorithms are tested under different scenario. (a) Urban macro; (b) Suburban macro; (c) Urban micro. The results are averaged by $50$ independent realizations.}}
\caption{\blue{NMSE versus the iteration number of various algorithms under urban macro scenario at SNR $= 10$ dB, with $N=256$ and $\frac{M}{N}=$ 0.4 in (a), $\frac{M}{N}=$ 0.6 in (b), and $\frac{M}{N}=$ 0.8 in (c). The results are averaged by $50$ independent realizations.}}
\caption{Experimental snapshots of a dissolving water/ethanol (v/v 40:60) drop in anethole oil experiencing different characteristic states during the dissolution. The first column of photos are top views and the second column are the corresponding side views. The initial drop size is around $\SI{0.5}{\micro\liter}$ and $t_0$ denotes the drop deposition time. (A) The drop begins with a transparent appearance surrounded by a clean host liquid. The arrow indicates solute plume above the drop. (B) W-in-o emulsion (water microdroplets) suspends outside the drop. The arrows and the inserted zoom-in panel highlights the location of the microdroplets. %Also inside droplet nucleation can be observed, see arrows. (C) O-in-w emulsion (oil microdroplets) appears inside the drop with a preferential location around the equator of the drop (arrows, area in between the two horizontal lines). (D) More oil microdroplets are formed and concentrate at the two sides of the equator. The w-in-o emulsion disappears. (E) In around 7 minutes, the two concentrated clusters of microdroplets evolve into two rings. The scale bar is \SI{0.5}{\milli\meter} in all those figures. \textcolor{blue}{(Corresponding movie: supplementary material Movie S2 and Movie S3)} }
\caption{Experimental snapshots showing appearance or absence of emulsifications during the dissolution of water/ethanol drops with different initial water-to-ethanol volume ratios (A-E for 30:70, 40:60, 50:50, 60:40, and 70:30, respectively). %Experimental snapshots of the w-in-o emulsification (A) and the o-in-w emulsification (B) of five different dissolving water/ethanol drops with different initial water-to-ethanol ratios (v/v 30:70, 40:60, 50:50, 60:40, 70:30). The first column of the photographs show that water-in-oil emulsification only happens for the high initial ethanol concentration cases (\SI{60}{vol\percent} and \SI{70}{vol\percent}). The water microdroplets around drops are pointed at by black arrows in the zoom-in figures of D and E. The last two columns of the synchronized side- and top-view photographs display the oil-in-water emulsification occurs inside drops in all the cases. The o-in-w emulsion are pointed at by the arrows. %The side-view photographs in A show that water-in-oil emulsification only happens for the high initial ethanol concentration cases (\SI{60}{vol\percent} and \SI{70}{vol\percent}). %The synchronized side- and top-view photographs in B display o-in-w emulsification occurring in all the cases. The scale bars are \SI{0.5}{\milli\meter} in all these figures. Some experimental videos are available \textcolor{blue}{in the supplementary material Movie S4}. }
\caption{Example of crops queried on a $4 \times 3$ grid for semantic segmentation. Some parts of the image (outlined in \textcolor{red}{red}) can be more interesting than others (outlined in {\color{blue}blue}). Querying labels for crops leads to a more effective utilization of the annotation budget than querying for images (best viewed in color).}
\caption{Layer configuration of ADNet model. The ADNet model is inspired from the VGG19 model -- the distinct additional layers in ADNet are indicated in {\color{blue}{blue}} color. The convolutional layers are indicated as conv\{receptive field size\}--\{number of channels\}. The fully connected layers are indicated as FC-\{number of channels\}. The dropout layer is indicated as Dropout (\{rate of dropout\}).}
\caption{Illustration of proposed variational models as a directed graph. \textit{(a)} VNLG: joint learning both variational parameters $\phi$ and generative model parameters $\theta$. \textit{(b)} DualVAE: \textcolor{red}{red} and \textcolor{blue}{blue} arrows form a standard VAE (parameterized by $\phi'$ and $\theta'$) as an auxiliary auto-encoding to the VNLG model denoted by \textcolor{red}{red} and black arrows.}
\caption{Comparison of top \textbf{Tv} responses generated for different models in different scenarios. Errors are marked in colors (\textcolor{red}{[missing]}, \textcolor{violet}{misplaced}, \textcolor{orange}{redundant}, \textcolor{teal}{wrong}, \textcolor{olive}{spelling mistake} information). \textcolor{blue}{[OK]} denotes successful generation. Model-X where X is amount of training data, \textit{i.e}. 10\%, 30\%, or 100\%.}
\caption{Comparison of top \textbf{Laptop} responses generated for different models in different scenarios. Errors are marked in colors (\textcolor{red}{[missing]}, \textcolor{violet}{misplaced}, \textcolor{orange}{redundant}, \textcolor{teal}{wrong}, \textcolor{olive}{spelling} information). \textcolor{blue}{[OK]} denotes successful generation. Model-X where X is amount of training data, \textit{i.e}. 10\%, 30\%, or 100\%.}
\caption{{Left}: First time conidial discharge intensity reaches given threshold according to the model, for different thresholds. Some trajectories did not reach given thresholds and have been omitted from the corresponding boxplots. {Right}: Duration that conidial discharge intensity is above given threshold. \hfill \small \textcolor{blue}{Blue}: 12.0 \gradc, \textcolor{green}{Green}: 18.5 \gradc, \textcolor{red}{Red}: 25 \gradc.}
\caption{Burning a graph in three rounds using a schedule defined by burning sequence $\langle A,B,C \rangle$. The number \blue{on each vertex} indicates the rounds at which \blue{the vertex becomes a burning vertex}. At round 1, a fire starts at $A$. At round 2, another fire starts at $B$ while the fire at $A$ spreads to all neighbors of $A$. At round 3, the fire spreads to all vertices except for $C$, where a new fire is started.}
\caption{Transformation of multi-colored graph to single-colored graph. Node $b$ can emit either \protect\bluecircle or \protect\redsquare, so it is split into nodes $bb$ and $br$, respectively. (This paper will use both color and shape to distinguish node ``colors.'')}
\caption{Steady-state (record length $\gamma=50$) tracking accuracy for the basic (black dashed), trackable (green dot-dash) and semi-unifilar (purple dotted) graphs. This is essentially a vertical slice of the data shown in Figure~\ref{fig:steadyStateContour}. The base graph starts around 77\% accuracy and never improves even for very long lags. The trackable graph has 100\% accuracy for lags $\beta\geq4$, but drops to the base 77\% level for shorter lags because of the same-colored out-neighbors of the form (\protect\bluecircle, \protect\redsquare). The semi-unifilar graph has 100\% accuracy for the lags shown, including real-time tracking ($\beta=0$).}
\caption{Accuracy $\alpha=P(\hat{X}_{t-\beta}=X_{t-\beta})$ as a function of lag $\beta$ and record length $\gamma$ for the (a) semi-unifilar and (b) observable graphs when observations start while the agent is at either a \protect\greendiamond or \protect\orangetriangle node. The white regions correspond to $\beta \geq \gamma$, for which the behavior is undefined.}
\caption{\label{fig:Fig_3} (Color online) Contour plots of the streamfunction $\sigma_{v}(x,y)$ and gauge invariant phase $\phi_{v}(x,y)$ within the superconducting annulus (\textcolor{mycolor}{inner radius} $\rho_{_0}=0.5$) when a vortex is present. Upper panel: the position of the vortex is given by $\rho_{v}=0.75$ and decreasing values of $\alpha_{v}$. For the lower panel: $\alpha_{v}=0$ and increasing $\rho_{v}$. The colormap has been shifted for visual purposes. \textcolor{mycolor}{Radial coordinates are given in units of $R$, i.e.: $\rho=r/R$}.}
\caption[3]{\label{Fig_4} Gauge invariant phase variation along the banks of the junction for different positions of the vortex given by the coordinates $(\rho_v , \alpha_v)$. \textcolor{mycolor}{As in Fig.\ref{fig:Fig_3}, here} $\rho_{_0}=0.5$. \textcolor{mycolor}{Radial coordinates are relative to the outer radius $\rho\equiv r/R.$} The inset illustrates the limiting situation in which the vortex is very close to the junction.}
\caption[4]{\label{Fig_5} \textcolor{mycolor}{Averaged phase difference between the two banks of the junction in terms of the radial position of a distant vortex ($\rho_v$) opposite to the aperture $(\alpha_{v}=0)$. The different lines correspond to superconducting loops of different widths given by the value of the inner radius. For each width, the vortex covers the range $\rho_{_0}<\rho_{v}<1$, i.e.: $R-W<r_v<R$). The linear approximation (dashed lines corresponding to Eq.(\ref{eq:linapprox})) is shown for the thinner loops.}}
\caption[4]{\label{Fig_6} Critical current pattern in terms of the applied magnetic field with no vortex present. The field is given in dimensionless units $\beta\equiv 2\pi W^2 B_a /\Phi_{_0}$. We plot the results for two different loops with respective normalized widths ${W/R}=0.75$ and ${W/R}=0.25$. \textcolor{mycolor}{The insets to the right show a detail of the streamlines of the current density ${\bf g}$. To the left, we show the dependence of the radial component $g_{\rho}^{+}$ close to the upper bank of the junction.}}
\caption{\label{fig:Fig_7} (Color online) Critical current patterns as a function of the applied magnetic field $\beta = 2\pi W^2 B_a /\Phi_{_0}$ in the presence of an antivortex close to the junction in the superconducting loop. The same loops as in Fig.\ref{Fig_6} are considered. The positions of the vortices (revealed by the streamlines $\sigma_{_{\rm L}}+\sigma_v = constant$ in the insets) are $\rho_v = 0.3, 0.625, 0.95$ and $\rho_v =0.8, 0.875, 0.95$ in units of $R$ respectively. \textcolor{mycolor}{Normalized units for $I_c$ are defined in terms of the zero-field, no-vortex value $I_c(0)$.}}
\caption[1]{\label{Fig_1}(Color online) The superconducting loop with a Josephson junction at the left. \textcolor{mycolor}{Schematically shown is the composition of conformal mapping transformations that convert the open annulus into the upper half-plane. Just for visual purposes, the width of the junction is oversized in the picture.}}
\caption{The thick line corresponds to Eq. (\ref{condition}), with $\gamma=1$ \magenta{and also $K=1$}. The dashed line is simply $H_i=M_D$. The dotted line is $H_i/M_D=M_D/\sqrt{B\gamma}M_P$ (or $T_{\mathrm{max}}=T_D$, see the texts). \magenta{The lower shaded region corresponds to $H_i<1$TeV, shown for } illustration. }
\caption{Examples of labeled data for training the scheduler network. \textcolor{red}{Red} and \textcolor{green}{green} boxes denote groundtruth and tracked results, respectively. (a) Positive examples, where the IOU of each groundtruth box and its corresponding tracked box is over a threshold; (b) Negative examples, where at least one such IOU is below a threshold or there are emerging/disappearing objects.}
\caption{Qualitative results of the scheduler network. \textcolor{red}{Red}, \textcolor{blue}{blue} and \textcolor{green}{green} boxes denote groundtruth, detected boxes and tracked boxes, respectively. The first row: R-FCN is applied in the keyframe. The second row: the scheduler determines to \textit{track} since it is confident. The third row: the scheduler predicts to \textit{track} in the first image although the red panda moves; however, the scheduler determines to \textit{detect} in the second image as the cat moves significantly and is unable to be tracked.}
\caption{\label{fig_higher_order}%\it\small Maximum nodal errors of Lagrange finite elements of degree $r=1,2,3$ on uniform anisotropic triangulations of types A and C in the domain $(0,2\eps)\times(0,1)$ for $u=e^{-x/\eps}$, $\eps=10^{-3}$ (left) and $\eps=1$ (right); the dashed lines correspond to $N^{-p}$ for $p=1,2,3,4$, {\color{blue}where $N$ and $\frac14N$ nodes are used in the $x$- and $y$-directions}.}
\caption{\label{fig_ABC}%\it\small Triangulations %of types A, B and C are obtained from the same rectangular tensor-product grid {\color{blue}(to be specified below)} by drawing diagonals in a different manner; the number of intervals in each coordinate direction is ${\mathcal O}(N)$.}
\caption{\label{Laplace_table_equi}\color{blue}Laplace equation with the exact solution $u=e^{-x/\eps}$ in the domain $(0,2\eps)\times(0,1)$, lumped-mass linear elements with $\{x_i\}$ uniform under the 1d Hessian metric: %$M=\frac14 N$: maximum nodal errors and computational rates $p$ in $N^{-p}$ }
\caption{SVM results on the test set for each of the 3 runs ($\star$), using the AUC score. The \textcolor{green}{\textbullet} and \textcolor{yellow}{\textbullet} represent the average AUC score of the 3 runs in both train and test sets respectively.}
\caption{Results for the \textit{Bulbul} CNN using the AUC score, for each of the 3 runs ($\star$). The \textcolor{green}{\textbullet} and \textcolor{yellow}{\textbullet} represent the average AUC score of the 3 runs in both train and test sets respectively.}
\caption{Classification pipeline during inference: \protect\tikz[inner sep=.25ex,baseline=-.75ex] \protect\node[circle,draw]{1}; and \protect\tikz[inner sep=.25ex,baseline=-.75ex] \protect\node[circle,draw]{2}; CNN Inference on the image, \protect\tikz[inner sep=.25ex,baseline=-.75ex] \protect\node[circle,draw]{3a}; Use the condition probabilities as a prior of the QA model, \protect\tikz[inner sep=.25ex,baseline=-.75ex] \protect\node[circle,draw]{3b}; Answer the QA questions based on the image description and \protect\tikz[inner sep=.25ex,baseline=-.75ex] \protect\node[circle,draw]{4}; compute the final condition probabilities }
\caption{\label{fig2}(Color online) A tiny position displacement of $\SI{4}{\angstrom}$ of the trapped ion is amplified to $\SI{10}{\nano\meter}$. The dashed lines indicate the probability distributions of the two weakly displaced wavepackets correlated with the eigenstates of $\hat{\sigma}_x$ in {\color{blue}{Eq.\,(4)}} with $g\simeq0.04$. The histogram shows the probability distribution of the finally postselected wavepacket reconstructed using the experimental data with $\theta=0.02$ and the solid line gives the exact theoretical prediction using {\color{blue}{Eq.\,(5)}}. }
\caption{(Color online) Uniformity calibration of coupling strengths for the red sideband and blue sideband lasers. The red line is a plot of {\color{blue}{Eq.\,(2)}} with parameters $\eta =0.08$ and $\Omega = 2\pi \times \SI{150}{\kilo \hertz}$. The black dots indicate the experimental measurement of $\ket{\uparrow}$ at chosen times. }
\caption{{\bf Schmidt values and overlaps of boundary MPS}. $\bfm{a.}$ shows the decay of the Schmidt values for boundary MPS representing all but the last layer of the network, cf. Fig.~\ref{fig:overlap}, split in the middle. Results for chimera graph with $N=2048$, results of which are shown in Fig.~{\color{blue} 3} of the main text. The spectrum is quickly decaying with growing $\beta$. $\bfm{b.}$ The respective overlap per site $O^{1/L}$, cf. Fig.~\ref{fig:overlap}. It is decaying with growing $\beta$ indicating on ill-conditioning of the problem. Collapse of curves for different linear system sizes $L$ points out that $O$ is vanishing exponentially with $L$. The plot shows median of 100 {\it droplet instances} with the error bars corresponding to 1-sigma of the distribution. $L$ is defined here as length of boundary MPS used to contract the network. Results obtained after employing the preconditioning procedure of Sec.~\ref{precond}. }
\caption{\red{Normalized probability distribution functions of the fractional uncertainties that $f_{\rm bol}$, \teff, and Distances from {\it Gaia\/} DR2 that contribute to the error on the calculated stellar radius. These PDFs have been generated from Kernel Density estimators that were fitted to the distributions for the final sample of single stars. Here we see that the typical contributions to the uncertainty on $R_*$ are $\lesssim 1\%$, $\lesssim 4\%$, and $\lesssim 3\%$, for distance, \teff, and $f_{\rm bol}$, respectively.}}
\caption{Left plot: Plot of the rotation period and \teff\for the Hyades single stars (filled-triangles) and the Pleiades single stars (filled-circles). Right plot: Same as the left plot but now plotting Rossby number($R_{\rm N}$) and \teff. All points within the two panels are sized proportionally to their calculated value of radius inflation ($\Delta R$). The three red crosses in the right panel serve as reference baselines. All points in the two panels are colored according to the sigma significance of the $\Delta R$. }
\caption{\textbf{Schematic of Raman-assisted nonlinear processes and their experimental implementation.} (a) The target six-quanta process that exchanges four photons of a high-Q resonator (magenta) with two excitations of a transmon mode (green) and vice versa. (b) Energy level diagram of a high-Q storage resonator at frequency $\omega_a$ coupled to a transmon mode at frequency $\omega_b$ (called conversion mode). The first three transmon eigenstates (denoted by $\ket{g}$, $\ket{e}$ and $\ket{f}$) and the first five eigenstates of the storage resonator (denoted by $\ket{0}$ to $\ket{4}$) are considered. Starting in $\ket{g0}$, the system is prepared in $\ket{f0}$ by applying $\ket{g}\rightarrow\ket{e}$ and $\ket{e}\rightarrow\ket{f}$ Rabi pulses (green arrows). A pump at frequency $\omega_{p1}$ (blue) connects $\ket{f0}$ to a virtual \textcolor{\reviewcolor}{(non-energy-conserving)} state, represented by the dashed line, detuned from $\ket{e2}$ with a detuning $\Delta$. \textcolor{\reviewcolor}{This virtual state acts as an intermediate metastable excitation of the transmon.} A second pump at frequency $\omega_{p2}$ (brown) connects the virtual state to $\ket{g4}$, thus converting the two transmon excitations into four resonator excitations. (c) Frequencies of the pumps and the transitions involved in the scheme. (d) Schematic of the implementation. The high-Q storage mode is formed by an aluminum $\lambda/4$-type 3-dimensional superconducting resonator (magenta), which is dispersively coupled to the conversion transmon (green) and the tomography transmon (red). The two $\lambda/2$ stripline resonators coupled to the transmons are used for performing single-shot readout of the respective transmons. \label{fig:figure_1}}
\caption{\coloronline Schematic of a true bicorrelation event in which two prompt fission neutrons are detected in coincidence with their originating fission. The schematic used is a two-dimensional view through an arc of the detector array in the \mcnpxpolimi model. }
\caption{\coloronline Bicorrelation time-of-flight distribution for (a) experimental data and (b) \polimi simulations, and bicorrelation energy distribution for (c) experimental data and (d) \polimi simulations.}
\caption{\coloronline Bicorrelation (a,b) time-of-flight and (c,d) energy distributions from \polimi simulation showing cross-talk effects, displayed for detector pairs at (a,c) \degrees{15} and (b,d) \degrees{45}. The diagonal bands in each distribution include cross-talk events, which move farther from the identity line ($\Dto=\Dtt$ and $\Eo=\Et$) and decrease in magnitude as the angle between detectors increases. }
\caption{\coloronline (a) Relative bicorrelation count rate \Wtheta for events from \MeV{1} to \MeV{4}, normalized by integral. (b) Ratio between relative bicorrelation rate for each simulation to that from experiment. The gray region from \degrees{0} to \degrees{20} serves as a reminder that cross talk is significant over this range.}
\caption{\coloronline (a) Magnitude of the neutron emission anisotropy, \Asym, as a function of \Emin and (b) ratio between simulated results and measured data. The magnitude of anisotropy increases as the neutron population is limited to higher energies.}
\caption{\coloronline (a) Average neutron energy as a function of bicorrelation angle across a range of \MeV{1} to \MeV{4}. (b) Ratio between average simulated energy and average measured energy, demonstrating agreement within 3\% across all data. The gray region from \degrees{0} to \degrees{20} serves as a reminder that cross talk is significant over this range.}
\caption{\coloronline Average energy of neutrons detected in coincidence with a (a) \MeV{2} and (b) \MeV{3} neutron.}
\caption{\coloronline Average correlated neutron energy \Ejave for fixed energy \Ei for detector pairs at (a) \degrees{85}, (b) \degrees{135}, and (c) \degrees{175}, and (d) slope of least-squares fit to $\Ejave(\Ei)$ at angles \degrees{85} and higher.}
\caption{\blue{The two unrooted binary phylogenetic trees $S_k$ and $S_k'$ as well as the generator $G_k$ (and $G_2$) that are used to show that the upper bound given in Lemma~\ref{l:tbr-cs} is tight for a pair of subtree and chain reduced trees. Blocks $A$, $B$, and $C$ indicate three chains each of length 3. For example, for $k=4$, block $A$ contains taxa 34, 35, and 36, block $B$ contains taxa 37, 38, and 39, and block $C$ contains taxa 40, 41, and 42. The edge $\{u,v\}$ of $S_k$ is used to argue that $k= d_\TBR(S_k,S_k')$ in the proof of Lemma~\ref{l:TBRsc=k}.}}
\caption{The unrooted binary phylogenetic trees $S_k$ and $S_k'$ as well as the generator $G_k$ (and $G_4$) that are used to show that the upper bound given in Lemma~\ref{l:tbr-cs} is tight for a pair of subtree, chain, and cluster reduced trees. Blocks $A$ and $B$ indicate two chains each of length 3. For example, for $k=6$, block $A$ contains taxa 64, 65, and 66, and block $B$ contains taxa 67, 68, and 69. \blue{The edge $\{u',v'\}$ of $S_k$ is used to argue that $k= d_\TBR(S_k,S_k')$ in the proof of Lemma~\ref{l:TBRscc=k}.}}
\caption{Qualitative results for the ground plane normal estimation. \textcolor{red}{Red}: groundtruth. \textcolor{blue}{Blue}: GroundNet. \textcolor{yellow}{Yellow}: Marr SkipNet. We show that our proposed GroundNet consistently outperforms all other baselines.}
\caption{Qualitative results for the horizon line estimation. \textcolor{red}{Red}: groundtruth. \textcolor{green}{Green}: GroundNet (DSAC). \textcolor{cyan}{Cyan}: Perceptual. \textcolor{yellow}{Yellow}: DeepHorizon. \textcolor{gray}{Gray}: Zhai et al. We show that our proposed GroundNet consistently outperforms all other baselines.}
\caption{\label{fig:sdmix}Equilibrium fraction of cooperators in well-mixed populations interacting in the Snowdrift game, $x^\ast$ (dashed line) and $x^\ast_k$ (dash-dotted line) for $k=4$, as a function of the cost-to-benefit ratio $r$ compared to individual based simulations (\textcolor{blue}{\scriptsize$\blacksquare$}) for $N=10^4$ and selection strength $\omega=0.01$. \textbf{a} random sampling of interaction partners matches $x^\ast$ even close to $0$ or $1$ indicating that for $N=10^4$ stochastic fluctuations, which may result in the extinction or fixation of cooperators, are negligible. \textbf{b} modified sampling scheme to include the competitor among the interaction partners matches $x^\ast_k$ and is consistently lower than $x^\ast$. Above $r_c=(k-1)/(k+1)=3/5$ cooperators can no longer persist. }
\caption{\label{fig:sdstruct}Equilibrium fraction of cooperators in structured populations interacting in the Snowdrift game as a function of the cost-to-benefit ratio $r$ from individual based simulations (\textcolor{blue}{\scriptsize$\blacksquare$}) with selection strength $\omega=0.01$. As a reference the equilibrium of the replicator dynamics is shown for the standard sampling scheme, $x^\ast$ (dashed line), as well as for the modified sampling scheme $x^\ast_k$ (dash-dotted line) with $k=4$. \textbf{a} lattice population ($N=100\times100$) interacting with the four nearest neighbours ($k=4$). Compared to $x^\ast$, spatial structure turns out to be detrimental to cooperation for larger $r$, however, compared to $x^\ast_k$ spatial structure remains beneficial. \textbf{b} random regular graph with four neighbours ($N=10^4, k=4$) results in equilibrium frequencies of cooperators between the lattice and $x^\ast$ due to the decrease in ordered structure and hence local clustering. }
\caption{\label{fig:sdbddb}Equilibrium fraction of cooperators in the spatial Moran process for the Snowdrift game as a function of the cost-to-benefit ratio $r$ from individual based simulations (\textcolor{blue}{\small$\bullet$}, standard deviation as error bars) for \textbf{a} death-birth updating and \textbf{b} birth-death updating with selection strength $\omega=0.01$ on a $N=100\times100$ lattice with four nearest neighbours ($k=4$). The simulation results are averaged over $100$ runs each, relaxed over $5000$ generations and then taking the mean and standard deviation over the next $5000$ generations. The equilibria for the Moran process in well-mixed populations in the limit $N\to\infty$ are the same as for the replicator dynamics for standard sampling (grey dashed line). Accounting for the fact that competitors also interact for birth-death updating recovers modified sampling (grey dash-dotted line). For the spatial Moran process the equilibria are derived from pair approximation in the limit of weak selection, $\omega\ll1$, for birth-death updating (black dashed line) and death-birth updating (dash-dotted line). }
\caption{\label{fig:shss}Stochastic stability in the Stag-hunt game for random matching in finite populations ($N=100$) for \textbf{\textsf{a}} standard sampling and \textbf{\textsf{b}} modified sampling (focal and model individuals always interact) with $k=4$ interaction partners and $\epsilon=0.001$ as a function of the payoff for a hare, $a$, see \eq{sh} (\textcolor{blue}{\small$\bullet$}, mean of stationary distribution with standard deviation as error bars). \textbf{\textsf{a}} for standard sampling, the risk dominant strategy is stochastically stable: for $a<1/2$ cooperation is risk dominant whereas for $a>1/2$ defection is risk dominant. \textbf{\textsf{b}} In contrast, for the modified sampling, cooperation is stochastically stable only for $a<1/4$, whereas defection remains the only stable state for $a>1/2$. However, for $1/4<a<1/2$ both homogeneous states are stochastically stable and attract the same probability mass, which results in a mean of $1/2$.}
\caption{\label{fig:sdss}Stochastic stability in the Snowdrift game for random matching in finite populations ($N=100$) for \textbf{\textsf{a}} standard sampling and \textbf{\textsf{b}} modified sampling (focal and model individuals always interact) with $k=4$ interaction partners and $\epsilon=0.001$ as a function of the cost-to-benefit-ratio, $r$, see \eq{sd} (\textcolor{blue}{\small$\bullet$}, mean of stationary distribution with standard deviation as error bars). \textbf{\textsf{a}} for standard sampling the interior equilibrium $x^\ast$ is always stochastically stable. Note, $x^\ast$ coincides with the homogeneous equilibria for $r=0$ and $r=1$. \textbf{\textsf{b}} In contrast, for the modified sampling, defection is stochastically stable for $r>3/5$, whereas the interior equilibrium remains stable for $r<3/5$. }
\caption{Performance of a three layer ConvNet with 32 filters each over $100$ epochs using \textcolor{orange}{standard scheme} and \textcolor{blue}{our method - RePr} on CIFAR-$10$. The shaded regions denote periods when only part of the network is trained. Left: Training Accuracy, Right: Test Accuracy. Annotations [A-F] are discussed in Section~\ref{sec:training}.}
\caption{Left: RePr training with various percentage of filters pruned. Shows average test accuracy over 5 epochs starting from epoch 20 for better visibility. Right:Marginal returns of multiple iterations of RePR - \textcolor{blue}{Training} and \textcolor{green}{Test} accuracy on CIFAR-10}
\caption{Test accuracy of a three layer ConvNet with 32 filters each over $100$ epochs using \textcolor{orange}{standard scheme} and \textcolor{blue}{our method - RePr} on CIFAR-$10$. The shaded regions denote periods when only part of the network is trained for RePr. Left: Fixed Learning Rate schedule of $0.1$, $0.01$ and $0.001$.Right: Cyclic Learning Rate with periodicity of $50$ Epochs, and amplitude of 0.005 and starting LR of 0.001.}
\caption{Test accuracy of a three layer ConvNet with 32 filters each over $100$ epochs using \textcolor{red}{standard scheme}, \textcolor{blue}{RePr with Oracle} and \textcolor{green}{RePr with Ortho} on CIFAR-$10$. Left: With Dropout of $0.5$. Right: No Dropout}
\caption{Localization performance comparisons for BUPM, BBS and DDIS. (Best viewed in digital version. Zoom in for details) Bounding box color: \textcolor{red}{\mysquare} BUPM, \textcolor{green}{\mysquare} BBS, \textcolor{blue}{\mysquare} DDIS}
\caption{AGILE-GRID and {\it Fermi}-LAT \gray{} fluxes and spectral indices. All dates are at 00:00:00 UTC.}
\caption{Multi-wavelength light-curves for the observing campaign on \src{}. {\bf Panel (a)}: GASP-WEBT 5\,GHz (black cross sign), 37\,GHz (blue triangles), 86\,GHz (red diamonds), and 228\,GHz(green squares) data[Jy]. {\bf Panels (b)--(h)}: $K$, $H$, $J$, $I$, $R$, $V$, $B$ bands (open circles, [mJy]). {\bf Panels (g)--(l)}: \sw{}/UVOT $v$, $b$, $u$, $w1$, $m2$, $w2$ bands (coloured discs, [mJy]). {\bf Panel (m)}: \sw{}/XRT observed 0.3--10\,keV flux[$10^{-11}$\,\ferg{}]. {\bf Panel (n)}: \agile{}/GRID (blue circles) and {\it Fermi}-LAT (red squares) data ($E>100$\,MeV,[$10^{-6}$\,\phcmsec]). The grey-dashed areas mark the time-interval (F1, MJD 57321.0--57327.0; F2, MJD 57333.0--57339.0) used for accumulating the almost simultaneous SEDs (orange (AGILE)/green (Fermi) and black(AGILE)/purple (Fermi) symbols, respectively) shown in Fig.~\ref{s0836:fig:sed}.}
\caption{In the first example, critical to answering the question correctly is discovering the presence of a fence (shown in \textcolor{red}{red}) in the attention heat-map. The cropping method of \cite{ChenBLL16} places a conservative box over this region, which corresponds to net-like or fence-like visual code-words like a tennis-net or a baseball batting-cage in the visual codebook. Similarly, in the second example, the attention corresponds to a visual code-word which clearly depicts boats, and in the third example, the attention corresponds to the teddy-bear code-word.}
\caption{We run a language-only \vqa baseline and note that although only 43\% of the questions are answered correctly in VQA 2.0 (\cite{GoyalKSBP17}), a large number of questions (88\%) in our experiments are answered with plausibly correct responses. For example, ``Sunglasses" would be a perfectly plausible answer to the question ``What does that girl have on her face?" - perhaps even more so than the ground-truth answer (``Nothing"). The \textcolor{red}{last example} shows an implausible answer provided by the model to the question.}
\caption{AID10: Matched tokens found in this message are (with the exception of final DID74 match) are signals that are likely correlated to the matched DID, but not an exact match. This is reflected in the significantly lower scores. Compare {\color{RedOrange} orange} and {\color{RoyalBlue} light blue} signals to those in \ref{fig:mp-top}. In \ref{fig:mp-bot} (top) the single bit mapped to DID73 seems to be a binary signal indicating whether the gas pedal is pressed. %Bytes 1-4 have been labeled as adjacent {\color{blue} dark blue} tokens with different endianness \textemdash inspection of plots show boundaries and labels are probably slightly off. }
\caption{gZSL learning results on CUB, AWA2 and aPY. ts = test classes (unseen classes), tr = train classes (seen classes), H = harmonical mean. The accuracy is class-average Top-1 in \%. The highest accuracy is in \textcolor{red}{red} color and the second is in \textcolor{blue}{blue} (better viewed in color).}
\caption{Zero shot learning results on CUB, AWA2 and aPY. SS = standard split, PS = proposed split. The results are class-average Top-1 accuracy in \%. The highest accuracy is in \textcolor{red}{red} color and the second is in \textcolor{blue}{blue} (better viewed in color).}
\caption{Top 10 header and body fonts used in header/body pairs. {\color{red} Ideally we should show them in corresponding fonts} }
\caption{The attention maps generated using LwM are more similar to `ideal' attention maps (i.e. maps generated with C), as compared to those generated with LwF-MC \cite{rebuffi2017icarl}. \red{[KC: We don’t need this figure if we already have a figure to illustrate the attention maps over time.]}}
\caption{\textcolor{verde}{\textit{Green}}}
\caption{\textcolor{arancione}{\textit{Orange}}}
\caption{\textcolor{rosso}{\textit{Red}}}
\caption{\textcolor{rosso}{\textit{Red}}}
\caption{An illustration of the difference between the exact analytical solution and the numerical solution for the SIA PDE. On the \textit{right} panel is a top view of the glacier, whose shape looks like a dome, and therefore the projection on to the x-y plane is a circle. The {\color{blue}blue points} signify the margin of the glacier (where it drops down to zero thickness), the {\color{red}red points} are at the interior of the glacier, and the {\color{black}black points} are towards the top of the glacier. The points that are not filled in signify the border of the glacier, where there is no ice thickness. On the \textit{left} panel the discrepancies between the analytical SIA PDE solution and the numerical SIA PDE solution for all grid points are shown. Specifically, the color of each path corresponds to the grid points on the right panel. Additionally, the paths are shown for 500 years, or 5000 time steps. }
\caption{Collecting the Jello (\textcolor{blue}{blue} box) requires more steps than the peach (\textcolor{orange}{orange} box) due to occluding objects. An object may be collected if it is within the magenta circle.}
\caption{Cluster error of models with different number of senses $S$. The vertical dashed lines correspond to the \textcolor{blue}{mean} and the \textcolor{red}{max} of the actual number of senses. The x-axes are log-scaled.}
\caption{The distribution of the height of the IIL (between zones 2 and 3) for a turbulent channel flow at \(Re_{\tau}=550\) (a) and \(950\) (b). The solid, red lines correspond to a \add{fitted log-normal probability distribution function}. (c) Mean streamwise velocity profile (normalized in wall units) is shown relative to the mean height of IIL between zone 2 and 3 for \(Re_{\tau}=950\) case. The log law \(U^+=(1/\kappa)\ln(y^+)+B\) (dashed line) is obtained by a linear fit of data between \(y^+=3Re_{\tau}^{1/2}\) and \(0.15Re_{\tau}\) (\(\kappa =0.41\)) following \citet{marusic_monty_hultmark_smits_2013}. The mean height is highlighted by (\protect\markertri) and (\protect\markercirc) corresponds to height of the quiescent core interface of \citet{kwon2014quiescent}. \add{The width of the vertical bar denotes the standard deviation of the interface height.}}
\caption{(a), (b), (e), (f) Reconstructed tomograms of a rabbit kitten head in agarose and (c), (d), (g), (h) rabbit kitten lungs in agarose with MICROFIL{\textregistered} contrast agent. Also shown are reconstructed slices of sample-free regions from both the (i), (j) head and (k), (l) lung data sets, containing only agarose. All data are shown without (first and third columns) and with (second and fourth columns) the correction detailed in this paper. Top row: Phase contrast tomograms of rabbit kitten head and lungs, no phase retrieval. Middle row: Tomograms of the same rabbit kitten head and lungs, after two-material phase retrieval \cite{Beltran2010, Croton2018}. Bottom row: Phase-retrieved tomograms of sample-free regions from rabbit kitten head and lung data sets. First and third columns: Standard dark-current and flat-field correction. Second and fourth columns: Beam sweep gain and offset correction.}
\caption{Illustration of common stages of image annotation: typically annotators first provide object class labels at the image-level~\cite{deng09cvpr,kuznetsova18arxiv} (\textcolor{red}{red}), sometimes associated to a specific object via a click as in~\cite{lin14eccv} and our approach (\textcolor{green}{green}). Following stages then annotate the spatial extent of objects, \eg with bounding boxes or segmentations (\textcolor{yellow}{yellow}).}
\caption{Example annotations on \ilsvrc{}. For each click we show the three alternatives from the ASR model (\textcolor{orange}{orange}) and the final class label (\textcolor{green}{green}). The first three images show typical annotations produced by our method. The last one shows a failure case: while the correct name is among the alternatives, an incorrect transcription matching a class name ranks higher, hence the final class label is wrong. % }
\caption{A comparison of typical mouse paths produced when annotating an image with our interface (\textcolor{darkgreen}{green}) or with~\cite{lin14eccv} (\textcolor{red}{red}). Circles indicate clicks. Mouse paths for our interface are extremely short, thanks to its simplicity and naturalness.}
\caption[]{\textbf{Visualization of latent representations.} The latent representations of corresponding \textcolor{greenforest_ours}{real (green; \ding{54})} and \textcolor{orangezinnia_ours}{synthetic (orange; \ding{58})} samples visualized using t-SNE~\cite{VanDerMaaten2008jmlr_tsne}. Visualization for a model trained on synthetic data (a) and our model trained on synthetic, unlabeled and 0.1\% of labeled real samples (b). Best viewed in color. }
\caption[]{\textbf{Visualization of latent representations.} t-SNE visualization of the learned latent representation of \textcolor{greenforest_ours}{real (green; \ding{54})} and \textcolor{orangezinnia_ours}{synthetic (orange; \ding{58})} samples from the validation set. Simultaneously, real and synthetic samples as well as similar poses are aligned in the latent representation, while only 100 corresponding real and synthetic images are employed during training. Note, if necessary, we moved the visualized depth images slightly apart, so, that they do not overlap. Best viewed in color with zoom. }
\caption[]{\textbf{Distributions of latent space distances between corresponding real and synthetic samples.} Comparison of the distance distributions for \textcolor{lightblue_ours}{our method} and a \textcolor{yellow_ours}{baseline}. The higher peak for lower distances shows that our method moves corresponding real and synthetic data closer together. See text for details.}
\caption{\textbf{Sketch of the architecture.} We train our system jointly with real and synthetic samples. From the joint latent representation $\mathbf{z}$, we predict the pose as well as two auxiliary outputs from which we also obtain feedback for unlabeled data. The auxiliary objectives are (i) to predict a different view and (ii) to discriminate between real and synthetic data. The training using unlabeled data ensures aligned latent feature distributions and thus, improved exploitation of synthetic data even with a small amount of labeled real samples. During test time only the pose is predicted (\textcolor{lightblue_ours}{blue path}). The layers per module are just for illustration and do not represent the actual number of layers. }
\caption{ A simple illustration of the differences between plan explicability, legibility and predictability. In this Gridworld, the agent can travel across cells, but cannot go backwards. Figure \ref{fig:1} illustrates a legible plan (\textcolor{green}{green}) in the presence of 3 possible goals of the agent, marked with {\bf ?}s. The \textcolor{red}{red} plan is not legible since all three goals are likely in its initial stages. In the parlance of transparent planning, the first action in the \textcolor{green}{green} plan can constitute a transparent plan (having conveyed the goal). Figure \ref{fig:2} illustrates an explicable plan (\textcolor{green}{green}) which goes straight to the goal {\bf G} as we would expect. The \textcolor{red}{red} plan may be more favorable to the agent due to its internal constraints (the arm sticking out might hit the wall), but is inexplicable (i.e. sub-optimal) in the observer's model. Finally, Figure \ref{fig:3} illustrates a predictable plan (\textcolor{green}{green}) since there is only one possible plan after it performs the first action. In the parlance of {\em t-predictability}, this is a 1-predictable plan. The \textcolor{red}{red} plans fail to disambiguate among two possible completions of the plan. Note that all the plans shown in Figure \ref{fig:3} are explicable (optimal in the observer's model) but only one of them is predictable -- i.e. explicable plans may not be predictable. Similarly, in Figure \ref{fig:2}, the \textcolor{red}{red} plan is predictable after the first action (even though not optimal, since there is only one likely completion) but not explicable -- i.e. predictable plans may not be explicable. Without a prefix in Figure \ref{fig:2}, the \textcolor{green}{green} plan is the only predictable plan. }
\caption{ A simple illustration of different goal obfuscation behaviors. Figure \ref{fig:2:1} shows different forms of deceptive behavior \protect\cite{masters2017deceptive} in \textcolor{red}{red} -- in simulation or ``hiding the truth'' (Figure \ref{fig:2:1a}) the agent could be going to either of the three possible goals while in dissimulation or ``showing the false'' (Figure \ref{fig:2:1b}) the likelihood of a decoy goal is strictly higher than that of the real goal. The \textcolor{green}{green} plan in (Figure \ref{fig:2:1a}) is a truthful plan. Figure \ref{fig:2:1} illustrates the difference between privacy and security of goal obfuscating plans. Here the observer cannot observe the actions of the agent in the first row of the grid due to occlusions. The \textcolor{red}{red} and \textcolor{green}{green} plans are both $2$-ambiguous \protect\cite{unified-anagha} and privacy preserving \protect\cite{keren2016privacy} -- the former allows for $\{G_1, G_2, G_3\}$ while the latter allows for $\{G_2, G_3\}$ as possible goal sets for the agent assumed to be rational. However a secure algorithm cannot flip from the \textcolor{red}{red} to the \textcolor{green}{green} plan when rerun with $G_2$. This is allowed under privacy preserving \protect\cite{keren2016privacy} and deceptive plans \protect\cite{masters2017deceptive} but not in secure plans \protect\cite{secure-anagha} -- i.e. the \textcolor{red}{red} plan is the only secure $2$-ambiguous solution. }
\caption{Boxplot of Dice score (in \textcolor{red}{red}) and IoU$_s$ (in \textcolor{blue}{blue}) per structure on CANDI-13 dataset. Only structures on the left hemisphere of the brain is shown for clarity. Center-lines indicate the median, boxes extend to the 25th and 75th percentiles, and the whiskers reach to the most extreme values not considered outliers (indicated by \textcolor{red}{red} crosses).}
\caption{(a) Comparison of the temperature before and after the Joule expansion at $\Delta = 1/2$ and $\lambda=1$. The filled symbols represent the temperatures obtained from the fitting of $P_{\rm L}$, while the open symbols from the isotherm curves. The dotted line represents the line where $T_f = T_i$. (b) Inversion curve at $\lambda=1$. The inversion temperature at $\Delta = 1/2$ is marked with the symbol {\color{orange} $\blacklozenge$}.}
\caption{\textbf{MonoGRNet for 3D object localization from a monocular RGB image.} MonoGRNet consists of four subnetworks for 2D detection(\textcolor[RGB]{237,125,49}{brown}), instance depth estimation(\textcolor[RGB]{112,173,71}{green}), 3D location estimation(\textcolor[RGB]{91,155,213}{blue}) and local corner regression(\textcolor[RGB]{255,192,0}{yellow}). Guided by the detected 2D bounding box, the network first estimates depth and 2D projection of the 3D box's center to obtain the global 3D location, and then regresses corner coordinates in local context. The final 3D bounding box is optimized in an end-to-end manner in the global context based on the estimated 3D location and local corners. (Best viewed in color.)}
\caption{Schematic representation of the VAE architecture presented in the text. The size of each layer is indicated by the value within brackets. The \textcolor{RoyalBlue}{blue rectangle $X$} represents the input layer, which is connected to a stack of two consecutive fully connected layers (black boxes). The last of the two black box is connected to two layers with four nodes each (\textcolor{Red}{red boxes}), representing the $\mu_z$ and $\sigma_z$ parameters of the encoder pdf $p(z|x)$. The \textcolor{Green}{green oval} represents the sampling operator, which returns a set of values for the 4-dimensional \textcolor{Red}{latent variables $z$}. These values are fed into the decoder, consisting of two consecutive hidden layers of 50 nodes each (black boxes). The last of the decoder hidden layer is connected to the \textcolor{RoyalBlue}{three output layers}, whose nodes correspond to the parameters of the predicted distribution in the initial 21-dimension space. The \textcolor{Magenta}{pink ovals} represent the computation of the two parts of the loss function: the KL loss and the reconstruction loss (see text). The \textcolor{Magenta}{computation of the KL} requires 8 additional learnable parameters ($\mu_p$ and $\sigma_p$, represented by the \textcolor{RedOrange}{orange boxes} on the top-left part of the figure), corresponding to the means and RMS of the four-dimensional Gaussian prior $p(z)$. The total loss in computed as described by the formula in the bottom-left black box (see Eq.~(\ref{eq:loss})). \label{fig:VAE_schematics}}
\caption{\textcolor{blue}{This table shows categorization result on Artist Scenes using different channels. Gabriel may check this table}}
\caption{Hertzsprung-Russell diagram with evolutionary tracks from \cite{marigo13} for various stellar masses. Different colors indicate the location in the diagram of the various stars. $\gamma$~Cru (\textcolor{blue}{\Large$\star$}) and $\mu$~Gem (\textcolor{green}{\Large$\star$}), while the filled circles represent the comparison stars for the earlier papers in this program. A typical error bar is shown in the upper left side of the figure.}
\caption{The DLP lightcrafter module \textregistered}
\caption{\label{fig:3C120} \blue Parsec-scale image of the jet in the active galaxy 3C~120 observed by the Very Long Baseline Array at 2~cm within the MOJAVE program. Top: total intensity contour plot overlayed by the fractional polarization distribution represented by the color. Bottom: the lowest contour of the total intensity distribution showing the shape of the jet overlayed by the contour plot of the linear polarization map. The sticks represent the electric vector position angle (EVPA). Black bullets show the positions of the jet components used in our analysis, with labels according to their IDs in Table~\ref{tab:sources}. The synthesized VLBA beam is shown at the half-power level in the bottom left corner.}
\caption{\label{tab:sources} List of source components selected for the analysis. MOJAVE ID and Component ID correspond to the MOJAVE catalog \blue \cite{MOJAVE_XIII, Lister-et-al-in-prep}. \black Components with the longest time coverage are indicated in bold.}
\caption{\label{fig:result1} Expected \blue in the case of no ULDM effect \black (gray band, 68\% CL) and observed (thick red line) $L$ as a function of the period $T$. }
\caption{Topographic map of the study area. Available stations with a minimum of 50\% observations during 1981 to 2013 are indicated by a red dot. The two boxes show the location of the Guinea Coast and West Sahel used in this study. Black triangles show prominent orographic features that are discussed in the text: 1.~Guinea Highlands, 2.~Jos Plateau, 3.~\blue{Cameroon Line}, 4.~Bongo Massif, 5.~\blue{Darfur Mountains}, and 6.~Ethiopian Highlands.}
\caption{Rainfall composite for days with significant ER wave signal\blue{, based on OLR anomalies} over \ang{5}-\ang{15}N, \ang{0}E (dashed line), for CHIRPS (\blue{shading}) and KASS-D (circles, both 1981--2013) \blue{during the extended monsoon season (April - October)}. As calculated with a bootstrap test, non-significant anomalies ($p>0.05$) are left white for the CHIRPS data, KASS-D stations that are non-significant are not shown. The number of days used for the composites and the fraction of days per season during each phase is given in the upper right of the plot. The numbers in the lower left part of the subpanels states the mean observed rainfall in all three rainfall datasets for the given phases (TRMM period: 1998--2013) within the Guinea Coast and West Sahel (black boxes). The modulation intensity, measured as the difference between the mean rainfall amount in the wettest (blue number) and driest phase (red number), is summarized at the right bottom.}
\caption{Mean precipitation in the eight phases in the Guinea Coast box (left side), and West Sahel box (right side) in CHIRPS (1981--2013, \blue{blue dotted line}), TRMM (1998--2013, red dashed line), and KASS-D (1981-2013, \blue{yellow solid line}) \blue{during the extended monsoon season (April - October). Because KASS-D lags the other datasets by \SI{6}{\hour}, the curve has been shifted using the mean period of each wave (in brackets). See section \ref{subsec:intens} for more detail}. Note that the y-axis has been scaled with the mean precipitation during all phases in the TRMM dataset, which is indicated by the horizontal red line.}
\caption{Modulation of rainfall by tropical waves in the Guinea Coast (\ang{5}W-\ang{5}E, \ang{5}-\ang{10}N) and the West Sahel (\ang{5}W-\ang{5}E, \ang{10}-\ang{15}N) during the transition season (Apr.--Jun.; Oct., top) and the full monsoon season (Jul.--Sep., bottom) \blue{as measured by rain gauges in KASS-D}. Dates in wet and dry phases of the tropical waves were obtained by filtering OLR in a Guinean (\ang{5}-\ang{10}N) and Sahelian band (\ang{10}-\ang{15}N) from 1979-2013. As a metric for the modulation intensity, the difference of the mean precipitation during wet phases and dry phases was calculated for all stations (Guinea n=106, Sahel n=117). From these stations an empirical probability density function was derived using kernel density estimation. Lines below the axis show mean and standard deviation of the corresponding modulation intensity for each wave. For abbreviations of wave names, see Table~\ref{tab.wave_characteristics}.}
\caption{Relative importance of tropical wave signals for TRMM precipitation (1998-2016) on different timescales over the Guinean band (\ang{5}-\ang{10}N, bottom panel) during the transition season (April to June and October). Explained variance is estimated by the squared correlations of the wave signal with raw precipitation. \blue{Lines are stacked on top of each other, such that the sum of all lines can be understood as the maximum variance explained by all wave types.} Significant correlations ($p<0.05$) are indicated with saturated colors. The black solid line shows the variance of precipitation at the specific timescale as percentage of the total variance in the raw data (3h). Band-averaged daily precipitation and surface height is given in the second lowest panel.}
\caption{(\textit{Best viewed in color}) Illustration of the entire framework of PCN. Image examples are sampled from our two-phase medical segmentation dataset (see Section~\ref{Experiments:Settings}). The learnable modules include two segmentation networks and a bidirectional generator. During training, there are three loss functions to compute, namely the {\em arterial}/{\em venous} segmentation losses and the generator loss. \textcolor{red}{Red} and \textcolor{blue}{blue} arrows indicate information propagation in the {\em arterial} and {\em venous} phases, respectively.}
\caption{\label{fig:contour} Contour plots for the spectral parameters measured from all observations (Observation one is \textcolor{red}{red}, observation 2 is \textcolor{blue}{blue} and observation 3 is \textcolor{yellow}{yellow}). The contours correspond to the $68\%$, $90\%$ and $99\%$ confidence levels.}
\caption{(a) We train the point cloud autoencoder to learn a latent space for 3D shapes (b) A set of 3D point clouds are mapped to the latent codes by the trained point cloud encoder. A GMM as a probabilistic model of the latent space is learned by fitting on this set of latent codes. (c) At inference time, we jointly estimate the shape and pose guided by the probabilistic shape prior via the learned GMM and the silhouette-shape constraint. The \textcolor{Orange}{\textbf{orange}}, \textcolor{SeaGreen}{\textbf{green}}, \textcolor{ProcessBlue}{\textbf{blue}}, and \textcolor{Fuchsia}{\textbf{purple}} blocks correspond to the components for shape, pose, silhouette, and loss terms respectively. The red arrows \textcolor{red}{$\bm{\longrightarrow}$} represent the gradient flow with respect to the latent code and the object pose during inference.}
\caption{The number of selected features ($\#$F) and the bootstrap classification accuracy (acc.) comparison for CMI-heuristic and CMI-permutation against different stopping criteria. We set $\alpha=1.01$ for R{\'e}nyi's $\alpha$-entropy functional. All criteria are ranked based on the difference between their selected number of features and the optimal values. The best two ranks are marked with \textcolor{darkgreen}{green} and \textcolor{blue}{blue} respectively. The average rank across all datasets is reported in the bottom line. The value behind the name of each dataset indicates the total number of features.}
\caption{The number of selected features ($\#$F) and the bootstrap classification accuracy (acc.) comparison for CMI-heuristic and CMI-permutation against different stopping criteria. We set $\alpha=2$ for R{\'e}nyi's $\alpha$-entropy functional. All criteria are ranked based on the difference between their selected number of features and the optimal values. The best two ranks are marked with \textcolor{darkgreen}{green} and \textcolor{blue}{blue} respectively. The average rank across all datasets is reported in the bottom line. The value behind the name of each dataset indicates the total number of features.}
\caption{ {\color{red} [TO REMOVE?] \tb{Comparison example} The layout reuslt(left column), corner map(middle column) and boundary map (right column) output by LayoutNet. Form top to bottom row are official pre-train model, training on Realtor360 (4-only) and training on Realtor360 (all). } }
\caption{Classification results for human observers (\textcolor{human.100}{red circles}) and ImageNet-trained networks AlexNet (\textcolor{alexnet.100}{purple diamonds}), VGG-16 (\textcolor{vgg.100}{blue triangles}), GoogLeNet (\textcolor{googlenet.100}{turquoise circles}) and ResNet-50 (\textcolor{resnet.IN}{grey squares}). Shape vs. texture biases for stimuli with cue conflict (sorted by human shape bias). Within the responses that corresponded to either the correct texture or correct shape category, the fractions of texture and shape decisions are depicted in the main plot (averages visualised by vertical lines). On the right side, small barplots display the proportion of correct decisions (either texture or shape correctly recognised) as a fraction of all trials. Similar results for ResNet-152, DenseNet-121 and Squeezenet1\_1 are reported in the Appendix, Figure~\ref{fig:biases_oidv2_deep_wide}.}
\caption{Shape vs. texture biases for stimuli with a texture-shape cue conflict after training ResNet-50 on Stylized-ImageNet (\textcolor{resnet.SIN}{orange squares}) and on ImageNet (\textcolor{resnet.IN}{grey squares}). Plotting conventions and human data (\textcolor{human.100}{red circles}) for comparison are identical to Figure~\ref{fig:biases_before_training}. Similar results for other networks are reported in the Appendix, Figure~\ref{fig:biases_vgg_alexnet_after_training}.}
\caption{Classification results for human observers (\textcolor{human.100}{red circles}) and ImageNet-trained networks AlexNet (\textcolor{alexnet.100}{purple diamonds}), VGG-16 (\textcolor{vgg.100}{blue triangles}), GoogLeNet (\textcolor{googlenet.100}{turquoise circles}) and ResNet-50 (\textcolor{resnet.IN}{grey squares}) on stimuli with a texture-shape cue conflict generated with style transfer, and \emph{biased} rather than neutral instructions to human observers. Plotting conventions and CNN data as in Figure~\ref{fig:biases_before_training}.}
\caption{Texture vs shape biases on of AlexNet and VGG-16 after training on Stylized-ImageNet. Plotting conventions as in Figures~\ref{fig:biases_before_training} and \ref{fig:biases_after_training}. Plot shows biases for AlexNet (\textcolor{alexnet.100}{purple diamonds}), VGG-16 (\textcolor{vgg.100}{blue triangles}) and human observers (\textcolor{human.100}{red circles}) for comparison. For GoogLeNet, no data is available since network training was performed in PyTorch and \texttt{torchvision.models} unfortunately does not provide a GoogLeNet (inception\_v1) architecture.}
\caption{Classification results for human observers and CNNs on stimuli with a texture-silhouette cue conflict (filled silhouette experiment). Plotting conventions as in Figures~\ref{fig:biases_before_training} and \ref{fig:biases_after_training}.\\ \textbf{Left:} Human observers (\textcolor{human.100}{red circles}) and ImageNet-trained networks AlexNet (\textcolor{alexnet.100}{purple diamonds}), VGG-16 (\textcolor{vgg.100}{blue triangles}), GoogLeNet (\textcolor{googlenet.100}{turquoise circles}) and ResNet-50 (\textcolor{resnet.IN}{grey squares}).\\ \textbf{Right:} Human observers (\textcolor{human.100}{red circles}, data identical to the left) and ResNet-50 trained on ImageNet (\textcolor{resnet.IN}{grey squares}) vs. ResNet-50 trained on Stylized-ImageNet (\textcolor{resnet.SIN}{orange squares}).}
\caption{The texture bias on cue conflict stimuli is not specific to ImageNet-trained networks (left) and also occurs in very deep, wide and compressed networks (right).\\ \textbf{Left:} The texture bias is not specific to ImageNet-trained networks. Comparison of texture-shape biases on cue conflict stimuli generated with style transfer for ResNet-101 trained on ImageNet (\textcolor{resnet.IN}{grey squares}) and ResNet-101 trained on the Open Images Dataset V2 (\textcolor{resnet101.oidv2}{green squares}) along with human data for comparison (\textcolor{human.100}{red circles}). Both networks have a qualitatively similar texture bias. We use a ResNet-101 architecture here since Open Images has released a pre-trained ResNet-101.\\ \textbf{Right:} The texture bias also appears in a very deep network (ResNet-152, \textcolor{resnet.IN}{grey squares}), a very wide one (DenseNet-121, \textcolor{densenet121.IN}{purple trianlges}), and a very compact one (SqueezeNet1\_1,\textcolor{squeezenet.IN}{brown diamonds}). Human data for comparison (\textcolor{human.100}{red circles}). All networks are pre-trained on ImageNet.}
\caption{Each lap is plotted with the colors of the learner whose action has been chosen: \textcolor{OliveGreen}{GPS \ac{NN}}, \textcolor{Orange}{Left camera \ac{NN}} and \textcolor{red}{Right camera \ac{NN}}. The algorithm was tested on the track 3 times for a total of 51 laps without failure of the task. }
\caption{Noisy EK-FAC. A subscript \(l\) denotes the index of a layer, \(\weights_l = \kvec(\mathbf{W}_l)\), and \(\mean_l = \kvec(\mathbf{M}_l)\). We assume zero momentum for simplicity. \(\oslash\) denotes element-wise division and \(\mathbf{r}_l^\gamma = \textrm{diag}(\mathbf{R}_l^{\gamma})\). unvec\((\cdot)\) is an inverse of vec\((\cdot)\) operation i.e. \(\textrm{vec}(\mathbf{A}) = \mathbf{a}\) and \(\textrm{unvec}(\mathbf{a}) = \mathbf{A}\). Differences from standard EK-FAC are shown in {\color{blue} blue}.}
\caption{Noisy K-FAC. A subscript \(l\) denotes the index of a layer, \(\weights_l = \kvec(\mathbf{W}_l)\), and \(\mean_l = \kvec(\mathbf{M}_l)\). We assume zero momentum for simplicity. Differences from standard K-FAC are shown in {\color{blue} blue}.}
\caption{A typical example of a planetary system evolving under the influence of neighbouring stars (Flammini et al., in prep.). {\it Left:} semi-major axis, $a$, and distance from the cluster centre, $r_c$. {\it Right:} eccentricity, $e$, and distance to the nearest neighbour star, $d_n$. The planetary systems were evolved using the \lonelyplanets{} code (\cite[Cai et al.2015, 2017; Flammini et al., in prep.]{cai2015,cai2017}), which combines the stellar dynamics (\nb; \cite{wang2016}) with that of the planetary systems (\rebound; \cite{rien2012}) in the \amuse{} framework (\cite{pelupessy2013}). We model star clusters with $N=500-10^4$ stars and radii $r_{\rm vir}=1$\,pc in a Solar-orbit tidal field. Stars are drawn from the Kroupa\,(2001) IMF, with masses of$0.1-25\,\msun$. Stars in the mass range $0.9-1.1\,\msun$ are assigned a planetary system similar to our own Solar system ({\color{blue}Earth}, {\color{green}Mars}, {\color{red}Jupiter}, {\color{cyan}Saturn}, {\color{violet}Uranus} and {\color{yellow}Neptune}). All models are integrated for 50\,Myr.}
\caption{\model: a network for guided segmentation of lung nodules, composed by a block responsible for predicting the initial segmentation and a second block for its correction. \protect\rotatebox{90}{$S$} is the side of the feature map. \protect\includegraphics[height=1.5ex]{figures/input_symbol}~input image {\color{red}$\blacksquare$} intermediary feature maps; {\color{orange}$\blacksquare$}~initial segmentation prediction; {\color{green}$\blacksquare$}~weight map \M{} computed from the user's input; {\color{blue}$\blacksquare$}~corrected segmentation. $\blacktriangleright$~$3\times 3 \times 3 \times N$ convolution, followed by batch normalization and rectified linear unit activation ($N$ is the number of feature maps, indicated on the top of each layer); $\blacktriangledown$~$3\times 3 \times 3 \times N$ convolution with stride $2\times 2 \times 2$, followed by batch normalization and rectified linear unit activation; $\blacktriangle$ $2 \times 2 \times 2$ nearest neighbor u-psample;$\triangleright$~$3\times 3 \times 3 \times N$ convolution with sigmoid activation. \label{fig:iwnet}}
\caption{Examples of weight maps (middle slice is shown) with different decay values $p$. Colorbar: 0~\protect\includegraphics[width=4em,height=0.5em]{colorbar}~1 \label{fig:weight_maps}}
\caption{Automatic and interactive lung nodule segmentations using \model{}.{\color{orange}$\blacksquare$}~ground-truth; {\color{red}$\blacksquare$}~prediction; {\color{green}$\blacksquare$}~end-points. \label{fig:pipeline}}
\caption{Examples of segmentations proposed by \model{}. For each of the $3\times 3$ \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_1} block: \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_2} ground-truth ({\color{red}$\blacksquare$}) and output of the first block of \model{} ({\color{orange}$\blacksquare$}) for two different annotators; \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_3} weight maps created based on the end-points of the diameter; \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_4} resulting segmentations after considering the diameter's end-points ({\color{green}$\blacksquare$}); \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_5} example of a 3D representation of the ground-truth from the nodule above; \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_6} 3D representation of the initial segmentation; \protect\includegraphics[width=0.8em,height=0.8em]{figures/image_location_7} 3D representation of the guided segmentation. \qquad\qquad \label{fig:seg_examples}}
\caption{Average Intersection over Union per nodule radius for the initial segmentation of \model{} ({\color{red}$\blacksquare$}) and the inter-observer agreement ({\color{yellow}$\blacksquare$}), and the respective standard deviation. \label{fig:rad}}
\caption{Average Intersection over Union per nodule texture for \model{}'s initial ({\color{red}$\blacksquare$}) and corrected segmentations ({\color{green}$\blacksquare$}), the inter-observer agreement ({\color{yellow}$\blacksquare$}), and the respective standard deviation.\label{fig:texture}}
\caption{Average absolute Intersection over Union improvement between the initial and the corrected segmentation using \model{} per nodule radius. Each column is normalized according to the respective number of nodules. %Only nodules with agreement level 2 and higher are considered. Colorbar: 0~\protect\includegraphics[width=4em,height=0.5em]{colorbar}~1.0\label{fig:iou_guided}}
\caption{Average surface distance (ASD) per nodule texture using \model{} for the initial segmentation ({\color{red}$\blacksquare$}), corrected segmentation ({\color{green}$\blacksquare$}) and the inter-observer agreement ({\color{yellow}$\blacksquare$}).\label{fig:asd}}
\caption{Some results of real time detection `\protect\includegraphics[scale=0.007]{Figure/a.jpg}',`\protect\includegraphics[scale=0.007]{Figure/ga.jpg}',`\protect\includegraphics[scale=0.007]{Figure/u.jpg}' \& `\protect\includegraphics[scale=0.0056]{Figure/dho.jpg}' using our proposed technique. This experiment had been done on various situation, i.e. different background, illumination and angles. In (a), `\protect\includegraphics[scale=0.007]{Figure/a.jpg}' has detection confidence rates of $98\%$ and $86\%$ (\textit{top} \&\textit{bottom row in left respectively}). For each row, $1$\textsuperscript{st} column shows the detection and the ($2$\textsuperscript{nd} column) has a zoom version of detected portion for better investigation. A reference sign is included at the rightmost column as ground truth collected from \textit{BdSLImSet}. Similarly, (b) shows that `\protect\includegraphics[scale=0.008]{Figure/ga.jpg}' has detection confidence rates of $90\%$ and $89\%$ (\textit{top} \&\textit{bottom}). In (c) and (d), `\protect\includegraphics[scale=0.008]{Figure/u.jpg}' has $94\%$ \&$99\%$ (\textit{top} \&\textit{bottom}), and `\protect\includegraphics[scale=0.007]{Figure/dho.jpg}' has $90\%$ \&$90\%$ confidence rate (\textit{top} \&\textit{bottom}) respectively.}
\caption{False recognition because of the similarity in the gestures of '\protect\includegraphics[scale=0.008]{Figure/a.jpg}' and '\protect\includegraphics[scale=0.008]{Figure/e.jpg}'. Here, for both rows, the $1$\textsuperscript{st} column shows the detection and the $2$\textsuperscript{nd} column has a zoom version of detected portion for better investigation. Both the rows show the same signs of '\protect\includegraphics[scale=0.008]{Figure/e.jpg}' while our approach performs perfectly in once case (\textit{top row in left}), while fails in another situation (\textit{bottom row}). A reference sign is included at the rightmost columns as ground truth.}
\caption{Bipartite graph with nodes from $\bar{X} = \{a_0,\ldots, a_3\} \cup \{\Delta_{b_0}, \Delta_{b_1}, \Delta_{b_2}\}$ labeled by \coloredcircle[blue, fill=blue]{3pt}'s and ${\color{red}\blacktriangle}$'s, and nodes from $\bar{Y} = \{b_0,\ldots,b_2 \} \cup \{\Delta_{a_0}, \ldots, \Delta_{a_3}\}$ labeled by \coloredbox{red} and ${\color{blue}\blacktriangle}$'s. The graph is complete only among off-diagonal points.}
\caption{(a): The predictive coding architecture used by the traditional video codec H.264 \cite{wiegand2003overview} or H.265 \cite{sullivan2012overview}. (b): The proposed end-to-end video compression network. The modules with \textcolor{blue}{blue color} are not included in the decoder side.}
\caption{Ablation study. We report the compression performance in the following settings. 1. The strategy of buffering previous frame is not adopted(\textcolor{red}{W/O update}). 2. Motion compensation network is removed (\textcolor{green}{W/O MC}). 3. Motion estimation module is not jointly optimized (\textcolor{blue}{W/O Joint Training}). 4. Motion compression network is removed (\textcolor{magenta}{W/O MVC}). 5. Without relying on motion information (\textcolor[rgb]{0.749,0.749,0.239}{W/O Motion Information}).}
\caption{Comparison of the average electrical frequency $f_\text{av}$ and the PFC power injection $P_{\text{PFC,tot}}$ obtained from the bulk ENTSO-E system ('\textcolor[rgb]{0.88627,0.03137,0.00000}{--}') and from the aggregated model \eqref{sweq}, \eqref{pfc} ('\textcolor[rgb]{0.74902,0.01176,0.38039}{- -}').}
\caption{Absolute and relative trajectories of the parameter estimates $\hat \eta_i,$ respectively $\frac{\hat \eta_i}{\eta_i}$ ($i=1$ '\textcolor[rgb]{0.88627,0.03137,0.00000}{--}' and $i=2$ '\textcolor[rgb]{0.74902,0.01176,0.38039}{- -}') with initial condition $\hat \eta(0)=\diag(0.3,0.2)\eta$.}
\caption{Trajectories of the parameter estimates $\hat \eta$ relative to the nominal parameter values $\eta$, {\em i.e.,} $\frac{\hat \eta_i}{\eta_i}$ ($i=1$ '\textcolor[rgb]{0.88627,0.03137,0.00000}{--}' and $i=2$ '\textcolor[rgb]{0.74902,0.01176,0.38039}{- -}') with initial condition $\hat \eta(0)=\diag(0.3,0.2)\eta$ in the rescheduling scenario.}
\caption{\label{fig:fail} \textbf{Failure example:} \small \it Input in \textcolor{red}{red}. \small \it InGAN has no semantic understanding of ``objects'' or ``scenes'', it only models the multiscale patch distribution of the input image, hence cannot distinguish between object-parts and entire objects. %Please see sec.~\ref{sec:failure_cases} for a detailed explanation. }
\caption{\label{fig:all_sizes} \textbf{InGAN's Capabilities:} \\small \it \emph{(Top:)} Once trained on an input image (marked by a {\color{red}{red frame}}), InGAN can synthesize a plethora of new images of significantly different sizes/shapes/aspect-ratios – all with the same ``DNA'' of the input image. All elements inside the image maintain their {local size/shape} and relative position. \textbf{Please view attached videos to see the continuum between different shapes/sizes/aspect-ratios}.\\\\emph{(Bottom:)} \InGAN provides a\emph{unified treatment for a variety of different datatypes} -- single/multi-texture images, painitings, and complex natural images, all under a single umbrella. }
\caption{\textbf{Ablation study:} \small \it Omitting $\Lrecon$ or using a single-scale $D$, degrades the results compared to full \textcolor{red}{InGAN} architecture.}
\caption[% Illustration of a von Mises-Fisher distribution and decision boundary of the cosine softmax classifier. ]{% \textcolor{red}{TODO: This could be combined with Figure 2 and made into a single column to save space.} Plot (a) illustrates a von Mises-Fisher distribution. The probability density increases as the cosine of the angle~$\alpha$ between sample~$\Feature$ and mean direction~$\tilde{\vec{w}}$ becomes smaller. Parameter~$\kappa$ controls the concentration of the distribution similar to how the standard deviation controls the spread of a Gaussian distribution. Plot (b) illustrates the posterior class probabilities (color coded) and decision boundary (white line) of the cosine softmax classifier for three classes. During training, all samples are pushed away from the decision boundary towards their parametrized class mean direction~(indicated by an arrow). }
\caption[Complexities and corresponding classes.]{Complexities and corresponding classes. “\tikz[baseline=-0.6ex]{\draw[->] (0, 0) -- (0.6, 0);}” denotes implication, “\tikz[baseline=-0.6ex]{\draw[->, dashed] (0, 0) -- (0.6, 0);}” denotes implication up to the equivalence. }
\caption{{\footnotesize \textcolor[rgb]{0.00,0.00,0.00}{Obtaining different variations of PTT values using PCG S1-peak and PPG different characteristic points. Corresponding PAT values are obtained with the use of ECG R-peak instead of PCG S1-peak as the proximal timing reference, where they are the same as PTT values plus a PEP offset.}}}
\caption{The mAP(\%) difference between target and translated features on three public datasets: Holidays (\textcolor[rgb]{0.6,0.8,0.5}{Green}), Oxford5k (\textcolor[rgb]{0.7,0.8,0.9}{Blue}) and Paris6k (\textcolor[rgb]{1.0,0.67,0.0}{Brown}) in the first, second and third blocks, respectively.}
\caption{The average mAP difference (\%) of MLP (\textcolor[rgb]{0.6,0.8,0.5}{Green}) and HAE (\textcolor[rgb]{0.7,0.8,0.9}{Blue}) on three datasets.}
\caption{Growth of \textcolor{blue}{$a(n)$} (odd partitions) against \textcolor{red}{$b(n+2)$} (chiral partitions) on a log-linear scale.}
\caption{ The input (\textcolor{red}{red}) and output (\textcolor{blue}{blue})}
\caption{% (a) Fraction of the total energy dissipation below a given distance from the wall, as a function of the Reynolds number. The dashed lines are scaled in wall units, and the labels refer to the predominant contributor to the dissipation in each region. The solid lines are three distances from the wall, scaled in outer units \citep[adapted from][]{jim18}. % \textcolor{refcolor2}{(b) Premultiplied spectra of \cite{hoy06} at $y^+=5$; \chndot, streamwise stress; \dashed, spanwise stress; \solid, pressure. The contours of each spectrum contain 50\%, 70\% and 90\% of its spectral mass. $\lambda_x$ and $\lambda_z$ are respectively the streamwise and spanwise wavelengths.} }
\caption{Parameters of the simulations. $L_{x,z}$ are the streamwise and spanwise period of the numerical box, and $N_{x,y,z}$ are the number of collocation points of the stored grid in each direction. \textcolor{refcolor2}{$\Delta_{x}^+,\Delta_{z}^+$ are the stream and spanwise grid resolutions of the DNS. $\Delta_{y_{\max(\min)}}^+$ is te maximum (minimum) resolution across the wall-normal direction. $Tu_\tau/h$ is the time spanned by the aproximatelly equispaced snapshots in eddy turnovers.} $D_y$ represents the discretisation used for the wall-normal direction: `CH' refers to Chebyshev polynomials and `FD' to compact finite differences. $N_f$ is the number of fields used to compute statistics.}
\caption{(a,b) Snapshots of S1000 at $y^+= 10$. (a) $v^+$. (b) $(v^\dagger)^+$. (c-h) F5300 at $y/h = 0.1$. (c) $u^+$. (d) $v^+$. (e) $(u^\dagger)^+$. (f) $(v^\dagger)^+$. \textcolor{refcolor1}{(g)} $u_G^+$, filtered with \r{eq:G} and $\Delta_x \times\Delta_z = (4\times 2)y$. \textcolor{refcolor1}{(h)} $v_G^+$ with $(2\times 2)y$. }
\caption{\textcolor{refcolor2}{Snapshot of the F5300 reconstruction. (a) The solid grey isosurfaces are $u^\dagger/u^{\dagger\prime}(y) < -1.5$, coloured with the distance from the wall, and the translucent red ones are $u_G/u_G^\prime(y) < -1.5$. (b) The solid grey isosurfaces are $u^\dagger v^\dagger/(u^\dagger v^\dagger)^\prime(y) < -1.5$, coloured with the distance from the wall, and the translucent red ones are $u_Gv_G/(u_Gv_G)^\prime(y) < -1.5$. The real velocities $u$ and $v$ (but not $u^\dagger$ and $v^\dagger$) are filtered using \eqref{eq:G} with $\Delta_x \times\Delta_z = (4\times 2)y$ and $\Delta_x \times\Delta_z = (2\times 2)y$, respectively.} }
\caption{Spectral reconstruction error at different heights and Reynolds numbers, on top of the premultiplied velocity spectra and cospectrum(shaded) for F5300. The line contours are $R_{ab} = 0.5$: \dashed F2000; \solid, F5300. The shaded contours contain $90\%, 50\%$ and $10\%$ of the total energy or of the tangential Reynolds stress. From left to right, $k_xk_z\phi_{uu}$, $k_xk_z\phi_{vv}$, $k_xk_z\phi_{ww}$ and $-k_xk_z\phi_{uv}$. From top to bottom $y^+\approx 20$, $y^+ \approx 40$, $y/h = 0.1$, $y/h = 0.2$. The blue-shaded portion of the spectra marks the region reproduced with less than $50\%$ error. }
\caption{Reconstructed energy ratio with single-observable models for F2000 (top) and F5300 (bottom). The lines are: \solid, ${a}^\dagger$; \chndot, ${a}^\dagger_{u_y}$; \dotted ${a}^\dagger_{w_y}$; \dashed, ${a}^\dagger_p$. From left to right, $uu$, $vv$, $ww$, $uv$. The dashed vertical lines are $y^+ \approx 100$, approximately corresponding to the end of the buffer layer.}
\caption{(a-d) Premultiplied spectra and cospectrum at $y/h = 0.1$ for F5300. Shaded, spectrum of the original fields; \solid, spectrum of the reconstruction using only the streamwise shear; \dashed, using only the pressure. The contour levels for all the spectra are 1.5\%, 10\% and 50\% of the maximum of the corresponding true spectrum. The white contours in (b) are reconstructions at $y/h = 0.05$. % The diagonal lines are: \solid, $\lambda_x=\lambda_z$; \dashed, $\lambda_x=6\lambda_z$. % (a) $k_xk_z\phi_{uu}$. (b) $k_xk_z\phi_{vv}$. (c) $k_xk_z\phi_{ww}$. (d) $-k_xk_z\phi_{uv}$. % (e-f) Premultiplied spectra of $v^\dagger_p$ (red) and $v^\dagger_{u_y}$ (black). (e) Modes along $\lambda_x = \lambda_z$. (f) Along $\lambda_x = 6\lambda_z$. % (g) Streamwise sections at different $\lambda_z\in (0.1-0.84)$, from dark to light, of: \solid, $k_xk_z\phi_{vv}^+ = 0.05$; \dashed, $k_xk_z\phi_{uu}^+ = 0.3$. % The straight lines are, \solid, $\lambda_x=\lambda_z$; \dashed, $\lambda_x=6\lambda_z$; \dotted, $y = 0.2\lambda_z$. }
\caption{Reconstruction of the wall-normal velocity at $y/h=0.1$ in F5300. Top, using only the pressure, $v^\dagger_p$. Middle, using only the wall shear, $v^\dagger_{u_y}$. \textcolor{refcolor2}{Bottom, true velocity field, filtered with \eqref{eq:G} and $\Delta_x \times \Delta_z (2\times 2)y$. The shading goes from $-u_\tau$ to $u_\tau$.} }
\caption{Snapshots of different reconstructions. (a) S2000, $v^+$ at $y^+\approx 10$. (b) S2000, $(v^\dagger_{\text{S1000}})^+$ at $y^+\approx 10$. (c) S2000 at $y/h \approx 0.1$. Shaded, $u^+$; \full, $(u^\dagger_{\text{S1000}})^+=1$; \dashed, $(u^\dagger_{\text{S1000}})^+=-1$. S2000 is filtered with $G$ and $\Delta_x \times\Delta_z = (0.4\times 0.2)$. % (d-f) S1000, $(u^\dagger)^+$ at $y/h \approx 0.1$, with information limited to discrete sensors (black crosses). (d) Fully dealiased. (e) No dealiasing. (f) Sensors averaged over 25\% of the wall surface. }
\caption{a) The conditions presented in Corollary \ref{cor:triangle-criterion} compare the weights of inner cuts (\textcolor{myorange}{- \negthinspace -}) and outer cuts (\textcolor{myblue}{- \negthinspace -}) around the triangle $\{u,v,w\}$. b) The conditions \eqref{eq:subgraph-multicut-condition} and \eqref{eq:subgraph-max-cut-condition}, presented in Theorem~\ref{thm:multicut-subgraph-criterion} and \ref{thm:max-cut-subgraph-criterion}, compare the weights of the inner cut $\delta(U, V_H \setminus U)$ and the outer cut $\delta(V_H) = \delta(U, V \setminus V_H) \cup \delta(V_H \setminus U, V \setminus V_H)$.}
\caption{Sensor ({\color{red}$\times$}) and actuator ({\color{blue}$\circ$}) placement for linearized Ginzburg-Landau. Each row corresponds to the optimized placement for a certain budget of sensors and actuators. Placements based on QR pivoting of balanced truncated modes~(a) closely approximate the $H_2$ norms of the placements determined using gradient minimization~(c). It is also possible to modify the QR algorithm to iteratively place sensors and actuators to avoid collocation (b). }
\caption{{\bf Dynamic State Changes.} {\bf (A)} State changes within trials in a {\bf (Ai)} random dot motion discrimination (RDMD) task, in which drift direction switches throughout the trial~\cite{glaze15}, and {\bf (Aii)} dynamic auditory clicks task, in which the side of the higher rate stream alternates during the trial~\cite{piet18}. {\bf (Aiii)} An ideal observer's LLR (See Eq.~(\ref{SDE}) in \textcolor{blue}{Box 1}) when the hazard rate is low (top panels: h=0.1Hz) and high (bottom panels: h=1Hz). Immediately after state changes, the belief typically does not match the state. {\bf (B)} State changes across trials. {\bf (Bi)} In the triangles task~\cite{glaze15}, samples (star) are drawn from one of two Gaussian distributions (yellow and red clouds) whose centers are represented by triangles. The observers must choose the current center (triangle). {\bf (Bii)} In an RDMD task, dots on each trial move in one of two directions (colored arrows) chosen according to a two-state Markov process. Depending on the switching rate, trial sequences may include excessive repetitions (Top), or alternations (Bottom). {\bf (Biii)} (Top) Responses can be biased by decisions from previous trials. (Bottom) Probabilistic feedback (`O': correct; `X': incorrect) affects initial bias (e.g., trials 3, 4, and 5), even when not completely reliable. }
\caption{Contours of normalised third octave band SPL across the Reynolds number range studied at $\alpha_g = 0^\circ$. The contours are normalised by their respective maximum SPL. The \protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=dor] (0.0,0.0) -- (0.15,0.0) (0.2,0.0) -- (0.35,0.0) (0.4,0.0) -- (0.55,0.0); \protect\end{tikzpicture} shows the $\sim Re_c^{1.5}$ trend line observed by \citet{Paterson1972}}
\caption{Comparison of the third octave sound pressure level (SPL$_{\protect\text{1/3}}$) at $Re_c$ = 300,000 between experimental result and the BPM prediction model. Where \protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=black] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture} is the experimental case and \protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=dor] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture} is the BPM model, which is a summation of: \protect\begin{tikzpicture} \protect\swrect{color=gray,fill=white}; \protect\end{tikzpicture} LBL, \protect\begin{tikzpicture} \protect\rect{color=gray,fill=white}; \protect\end{tikzpicture} pressure side TBL, \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[thick,color=gray,fill=white] (0.1,0.1) circle (0.1); \protect\end{tikzpicture} suction side TBL and \protect\begin{tikzpicture} \protect\triup{color=gray,fill=white}; \protect\end{tikzpicture} inflow noise.}
\caption{Single microphone measurements for the untripped cases. Each Reynolds number case, indicated on (a), is spaced by 30dB for clarity. Each of the angles stated are the geometric angles of attack. For each angle and Reynolds number there are three test cases: a baseline case with no flaplets (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=black] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}), the case where the flaplets are affixed to the pressure side (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=pinot] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}) and when the flaplets are affixed onto the suction side (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=royal] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}).}
\caption{Comparison of previous literatures and the present study in the tonal noise envelope for the NACA 0012 aerofoil as proposed by \citet{Lowson}. The angle of attack ($\alpha^*$) is corrected using the BPM empirical correction for the open jet wind tunnel results. Neither the direct numerical simulation (DNS) or closed wind tunnel angles have been adjusted. \protect\begin{tikzpicture} \protect\rect{color=darkgray,fill=white}; \protect\end{tikzpicture} \citet{Paterson1972}; \protect\begin{tikzpicture} \protect\swrect{color=darkgray,fill=white}; \protect\end{tikzpicture} \citet{Lowson}; \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[solid,very thick,color=darkgray] (0.1,0.0) -- (0.1,0.2) (0.0,0.1) -- (0.2,0.1); \protect\end{tikzpicture} \citet{Desquesnes2007}; \protect\begin{tikzpicture} \protect\tridown{color=darkgray,fill=white}; \protect\end{tikzpicture} \citet{Inasawa2013}; \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[thick,color=darkgray,fill=white] (0.1,0.1) circle (0.1); \protect\end{tikzpicture} \citet{Chong2013}; \protect\begin{tikzpicture} \protect\triup{color=darkgray,fill=white}; \protect\end{tikzpicture} \citet{Probsting2014}; \protect\begin{tikzpicture} \protect\tstar{0.05}{.13}{5}{54}{semithick,color=darkgray,fill=white}; \protect\end{tikzpicture} \citet{Arcondoulis2018}; \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[solid,thick,color=pinot] (0.0,0.0) -- (0.2,0.2) (0.0,0.2) -- (0.2,0.0); \protect\end{tikzpicture} Present (tonal) and \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[solid,thick,color=royal] (0.0,0.0) -- (0.2,0.2) (0.0,0.2) -- (0.2,0.0); \protect\end{tikzpicture} Present (non-tonal)}
\caption{Single microphone measurements for the tripped cases. Each Reynolds number case, indicated on \ref{fig: SingleMic-0deg}, is spaced by 30~dB for clarity. Each of the angles stated are the geometric angles of attack. For each angle and Reynolds number there are three test cases: a baseline case with no flaplets (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=black] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}), the case where the flaplets are affixed to the pressure side (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=pinot] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}) and when the flaplets are affixed onto the suction side (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=royal] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}).}
\caption{The spatial growth rate on the pressure side of the aerofoil at different chord wise positions (x/c) against the frequency where they occur. \protect\begin{tikzpicture} \protect\swrect{color=gray,fill=white}; \protect\end{tikzpicture} 0.4c, \protect\begin{tikzpicture} \protect\rect{color=gray,fill=white}; \protect\end{tikzpicture} 0.5c, \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[thick,color=gray,fill=white] (0.1,0.1) circle (0.1); \protect\end{tikzpicture} 0.6c, \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[solid,thick,color=black] (0.0,0.0) -- (0.2,0.2) (0.0,0.2) -- (0.2,0.0); \protect\end{tikzpicture} is the position of the laminar separation bubble; (a) 0.67c, (b) 0.72c, (c) 0.76c. \protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[dashed, thick,color=pinot] (0.0,0.0) -- (0.4,0.0); \protect\end{tikzpicture} Indicates the frequency where the tonal peak ($f_n$) is observed in the experiment, see fig.\ref{fig: SingleMic-10deg}.}
\caption{Overall sound pressure levels for the untripped and tripped baseline cases and the flaplet, pressure side mounted, cases. $\Delta$OSPL has been plotted on the second axis to yield a clear indication of the difference at each Reynolds number. Baseline untripped(\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=black] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}), untripped flaplets pressure side mounted (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=pinot] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}), tripped baseline (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=gray] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}) and tripped flaplets pressure side mounted (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[solid,very thick,color=royal] (0.0,0.0) -- (0.3,0.0); \protect\end{tikzpicture}). \protect\begin{tikzpicture} \protect\tikz[baseline=0pt] \protect\draw[thick,color=gray,fill=white] (0.1,0.1) circle (0.1); \protect\end{tikzpicture} Indicates the $\Delta$OSPL of the untripped case and \protect\begin{tikzpicture} \protect\rect{color=gray,fill=white}; \protect\end{tikzpicture} indicates the $\Delta$OSPL of the tripped case. The zero line on the $\Delta$OSPL axis is shown by (\protect\begin{tikzpicture} \protect\tikz[baseline=-2pt] \protect\draw[dashed, thick,color=gray] (0.0,0.0) -- (0.4,0.0); \protect\end{tikzpicture}).}
\caption{\label{Figure:Masses} (Color online) Typical examples of signal and background mass distributions from butanol data after applying all kinematic cuts and corrections. The invariant $\pi^+\pi^-\pi^0$ masses are shown for two energies at the same angle, $0.2\,<\,{\rm cos}\,\Theta^{\,\omega}_{\rm \,c.m.}\, < 0.4$. The {\bf black line} shows the unweighted distribution from the butanol target, the {\color{blue} blue-shaded area} shows the background mass distribution (weighted by $1-Q$), and the {\color{red} red line} shows the signal distribution (weighted by $Q$.)}
\caption{Definition of different viewpoints. % Viewpoints of one identity are sampled at an interval of 10$^\circ$. Left orientation represents the set of the viewpoints that contains more information on the left side of the person, \ie 320$^\circ$ - 40$^\circ$. Similarly, other orientations of the pedestrian represent the sets of viewpoints containing front (50$^\circ$ - 130$^\circ$), right (140$^\circ$ - 220$^\circ$) and back (230$^\circ$ - 310$^\circ$) information. The viewpoints with \textcolor{blue}{the blue tags} represent the due left (0$^\circ$), front (90$^\circ$), right (180$^\circ$) and back (270$^\circ$) of a person. }
\caption{Sample re-ID results on Market-1203. Images in the first column are queries. The retrieved images are sorted according to their similarity to the query (high to low) from left to right. The similarity is calculated by using feature extracted from the PCB model. True matches and false matches are in \textcolor{green}{green} and \textcolor{red}{red} rectangles, respectively. %\textbf{(a):} Results of the control group and ``remove 3 true matches of similar viewpoints to query" are displayed. %\textbf{(b):} For a query, the rank of its true matches is lower than some false matches with similar viewpoints. }
\caption{Nondimensionalized stiffness $k^{*}/\rho g$ and effective gravity $\tilde g/g$ over the width of the water waveguide $y/a$ for different values of $\delta$ ($\delta=0.7$ (\blue{\L}), $\delta=0.8 (\red{\cdashL})$, and $\delta=0.9$ ({\dashL})) where the waveguide's core effective gravity is considered to be very close to the gravitational acceleration $\tilde g/ g = 1.001$ }
\caption{\label{tab:interaction} \textbf{Future interaction prediction experiment:} Table comparing the performance of \method\with state-of-the-art algorithms, in terms of mean reciprocal rank (MRR) and recall@10. The{\color{blue!75}best algorithm} in each column is colored {\color{blue!75}blue} and {\color{blue!20}second best is light blue}. \method\outperforms the baselines by up to 22.4\%.}
\caption{\label{tab:churn} \textbf{User state change prediction:} Table comparing the performance in terms of AUC of \method\with state of the art algorithms. The{\color{blue!75}best algorithm} in each column is colored {\color{blue!75}blue} and {\color{blue!20}second best is light blue}. \method\outperforms the baselines by up to 4.5\%.}
\caption{\textbf{Robustness of \method :} Figures (a--c) compare the mean reciprocal rank (MRR) of \method\with baselines on interaction prediction task, by varying the training data size. Figure (d) shows the AUC of user state change prediction task by varying the training data size. In all cases,\method\is consistently the best by up to 33\%. \label{fig:interaction}}
\caption{\textbf{Robustness to dynamic embedding size:} Figure shows that the MRR of \method\is stable with change in dynamic embeddign size, for the task of interaction prediction on LastFM dataset. Please refer to the legend in Figure 5.\label{fig:embed} }
\caption{\textbf{\method\as an early-warning system:} Figure showing that as a student gets closer to dropping-out (moving right), \method\predicts a higher dropping out probability score for them compared to other students.\label{fig:ews}}
\caption{Top graph: the accuracy of \jitnet~0.8 and Offline Oracle relative to MRCNN. Bottom graph: the number of updates to \jitnet\during online distillation. Plotted points are averages over a 30~second interval of the video. Images correspond to circled points in bottom plot, and show times where\jitnet\required frequent training to maintain accuracy.}
\caption{The JITNet architecture. s = stride, r = output resize scale, c = output channels. \red{TODO: arrows, enc1 to dec1, enc2 to dec2}}
\caption{Results for FBMS, ComplexBackground (CB), CamouflagedAnimal (CA), and averaged over all videos in these datasets (ALL). Best results are highlighted in \textcolor{red}{red} with second best in \textcolor{blue}{blue}.}
\caption{Results on Video Foreground Segmentation for DAVIS2016 and FT3D. Best results are highlighted in \textcolor{red}{red}.}
\caption{Results on Video Foreground Segmentation for DAVIS2016 and FT3D. Best results are highlighted in \textcolor{red}{red}.}
\caption{Results on Motion Segmentation for FBMS, ComplexBackground (CB), and CamouflagedAnimal (CA), and averaged over all videos in these datasets. Best results are highlighted in \textcolor{red}{red} with second best in \textcolor{blue}{blue}.}
\caption{Overall architecture. First, feature maps of each frame are extracted from the Y-Net. Next, foreground masks are computed, shown in \textcolor{orange}{orange}. The PT-RNN uses these foreground masks to compute trajectory embeddings (example foreground trajectory from frame $1$ to $T$ shown in \textcolor{purple}{purple}), which are normalized to produce unit vectors. Backpropagation passes through the \textcolor{blue}{blue} solid arrows, but not through the \textcolor{red}{red} dashed arrows.}
\caption{We show U-Net \cite{ronneberger2015u} and our proposed Y-Net to visually demonstrate the difference. Y-Net has two encoding branches (shown in \textcolor{green}{green}) for each input modality, which is fused (shown in \textcolor{purple}{purple}) and passed to the decoder (shown in \textcolor{yellow}{yellow}). Skip connections are visualized as \textcolor{blue}{blue} arrows.}
\caption{Comparison of criticality index-based (black) and electricity yield-based (\textcolor{NavyBlue}{blue}) deployment of wind farms within a given subset of locations, based on the criticality indicator values. Example depicting the results for the subset of potential generation sites in France, $\mathcal L_F$.}
\caption{Table 1. Selection rules for second harmonic generation in nanoparticles with different shapes \blue{ (left columns), and examples of allowed multipoles on $2\omega$ in GaAs nanostructures under the $x$-polarized plane-wave excitation (right column).}}
\caption{Figure 2. Second-harmonic radiation patterns generated by a magnetic dipole (MD) at the fundamental frequency for the GaAs cylinder \blue{(top left)}, and the GaAs cylinder truncated laterally along the plane $x~=~y$ \blue{(bottom left)}. For the cylinder, the main generated mode is the magnetic quadrupole (MQ). In the truncated cylinder the electric dipole (ED) is also generated, which creates a radiation pattern directed along the $x = y$ axis when interfering with the magnetic quadrupole. \blue{The relative position of the $T_\text{d}$-lattice axes is shown}. The cylinder's radius is 140~nm, height is 280~nm, the wavelength is 1480~nm (for $\omega $) and 740~nm (for $ 2 \omega $), $ \varepsilon = 12.96 $. {The width} of the truncated cylinder is 230~nm. \blue{On the right side the illustration of derivation of the selection rules given by the integral \eqref{theI} is shown. Cartesian projections of vector spherical harmonics $[\vec N_{\alpha}\cdot \vec W ]$ are the scalar functions with particular symmetry which give invariant functions after multiplication.}}
\caption{Influence of geometry and interactions on the condensate fluctuations. (a) Maximal fluctuations of the number of condensed particles $\max( \Delta N_0^2 / N)$ as a function of the total number of atoms for the spherical trap (\mysquare{myblue}) and the traps in which experiments have been performed with aspect ratios $ 5.27 $ (\mybullet{red}) and $ 7.64 $ (\mydiamond{teal})~\cite{SupplementaryMaterial}. The horizontal gray line is the limiting value $\zeta (2) / \zeta (3)$ given in Eq.~(\ref{eq:Politzer}). (b) Fluctuations of the condensate atom number in the interacting system relative to the noninteracting system as a function of the interaction strength $n(0)^{1/3} a$, where $n(0)$ is the peak density and $ a $ the scattering length. (\mydiamond{myblue}) and (\mydiamond{red}) are results obtained within the classical field approximation for $N=100$ and $N=200$ atoms~\cite{SupplementaryMaterial}. The other points correspond to \mytriangle{teal}{60} \cite{Bhattacharyya2016}, \mysquare{purple} \cite{Idziaszek1999}, \mypentagon{grey}{36} \cite{Svidzinsky2006}. The vertical dashed line indicates a typical interaction strength in our experiments. }
\caption{Sample of SS3M phenotypes with their true labels (ICD9 diagnosis codes), expert generated labels, and qualitative evaluation results. \color{Red} Red tokens \color{Black} (top row): words from clinical notes. \color{Green} Green tokens \color{Black} (middle row): clinical labs. \color{Blue} Blue tokens \color{Black} (bottom row): medications. A token's size represents its probability relative to other tokens under the phenotype.}
\caption{\small AUC of \red{\(F_{cur}\)} vs. \blue{\(F_{cur+b}\)}. (a, b) show absolute and relative follower growth for increasing interval sizes (further in the future). (c) shows relative follower growth over 2-month intervals starting at different ages.}
\caption{\small AUC of \red{\(F_{cur}\)} vs. \blue{\(F_{cur+b}\)} over different interval sizes ($\delta$) to predict (left) average concurrent views, (middle) \# of cheers, and (right) cumulative views. Error bars denote 1 standard error.}
\caption{\small Coefficients of features in the $F_{cur+b}$ model for relative popularity growth over a two month time interval (*: p-value < 0.1, **: p-value<0.05). Third-party social media features are \gray{colored in gray}. Coefficients that changed from negative in the followers model to positive are colored \blue{blue}, and colored \red{red} if from positive to negative. }
\caption{ Evolution of the magnetic twist during 2012-06-14 07:00 to 14:00 UT. Panels on the left shows twist per unit length, while panels on the right give the total twist. From top to bottom are the results from three methods. Red (black) curves represent the values calculated for FP+ (FP-). \bluec{Errors here come from two sources: the uncertainty of calculation is estimated using errors proprogation; and uncertainty of measurement that related to the pixel selection ($B_{z} >20$ G, $B_{t} > 100$ G).} \label{fig5} }
\caption{{\color{blue} OPD distributions in aperture plane for subreflector adjustment. ($a$) pre-adjustment for 40° elevation angle, ($b$) after-adjustment for 40° elevation angle; ($c$) pre-adjustment for 70° elevation angle, ($d$) after-adjustment for 70° elevation angle.} }
\caption{\label{Up_fig} (Top) Inner-scaled tangential mean velocity profiles at $x_{ss}/c=0.6$ for the four wing cases under study. Colors as in Figure~\ref{beta_fig}, and {\color{mygray}\solid} denotes ZPG TBL data at matching $Re_{\tau}$. The matched $U^{+}_{t}$ profiles for W10 and ZPG10 are denoted by $\left ( \blacksquare \right )$, for W4 and ZPG4 by $\left ( \bullet \right )$ and for W2 and ZPG2 by $\left ( \blacklozenge \right )$. (Middle) Ratio of $U^{+}_{e}$ and (bottom) $H$ between wing and ZPG at matching $Re_{\tau}$, where (\textcolor{blue}{$\blacksquare$}), (\textcolor{green}{$\blacksquare$}) and (\textcolor{red}{$\blacksquare$}) denote ratios at $x_{ss}/c=0.4$, $0.6$ and $0.7$, respectively.}
\caption{\label{statistics_figure} (Left panel) Inner-scaled mean flow (with reference low-$Re$ values $\kappa=0.41$ and $B=5.2$), (middle panel) inner-scaled Reynolds-stress tensor components and (right panel) TKE budget scaled by $u_{\tau}^{4}/\nu$. Reynolds stresses are represented as: {\color{blue}\solid} tangential, {\color{red}\solid} wall-normal and {\color{green}\solid} spanwise velocity fluctuations, and {\color{cyan}\solid} Reynolds shear stress. Budget terms are represented as follows: {\color{blue}\solid} Production, {\color{red}\solid} Dissipation, {\color{green}\solid} Turbulent transport, {\color{cyan}\solid} Viscous diffusion, {\color{mygray}\solid} Velocity-pressure-gradient correlation and {\color{yellow}\solid} Convection. Data extracted at (top line) $x_{ss}/c = 0.8$ and (bottom line) $x_{ss}/c = 0.9$, and compared with the ZPG data by Schlatter and \"Orl\"u (2010).}
\caption{Nighttime localization results (values from~\cite{sattler:hal-01859660} accounting for the \textit{mixed-conditions} additional training images) for our approach and state-of-the-art feature-based and retrieval-based methods. (\textasteriskcentered) \textit{\textbf{NB:} AS + Sem. Match~\cite{Toft_2018_ECCV} leverages manually annotated nighttime to train its semantic segmentation network. ToDayGAN~\cite{DBLP:journals/corr/abs-1809-09767} uses 19,998 additional nighttime images to train its GAN network. In comparison, our method requires very few nighttime images to produce these results, and adds no overhead to the overall computational cost of retrieval-based methods.}}
\caption{\label{fig:graph2} Clauser pressure-gradient parameter $\beta$ along the suction side of the NACA4412 (\textcolor{blue}{---}) and the NACA0012 (\textcolor{red}{---}) wing sections.}
\caption{\label{fig:graph3} Streamwise evolution of friction Reynolds number. The TBLs developing on the suction side of the NACA4412 (\textcolor{blue}{---}) and the NACA0012 (\textcolor{red}{---}) wing sections are shown.}
\caption{\label{fig:graph4} Streamwise evolution of momentum-thickness Reynolds number. The TBLs developing on the suction side of the NACA4412 (\textcolor{blue}{---}) and the NACA0012 (\textcolor{red}{---}) wing sections are shown.}
\caption{Inner-scaled mean velocity profile along the wall-normal direction in the following cases: ZPG (\textcolor{green}{---}), NACA4412 (\textcolor{blue}{---}), NACA0012 (\textcolor{red}{---}). Note that the ZPG case was chosen to approximately match the corresponding $Re_\tau$ values of the wing sections.}
\caption{Selected inner-scaled Reynolds stresses: tangential (\textcolor{blue}{---}), normal (\textcolor{red}{---}), spanwise (\textcolor{green}{---}) velocity fluctuations and Reynolds-shear stress (\textcolor{black}{---}).}
\caption{Pre-multiplied spanwise spectra of the tangential velocity fluctuations for the NACA4412 (\textcolor{blue}{---}), NACA0012 (\textcolor{red}{---}) and ZPG at $Re_\theta = 880$ (\textcolor{green}{---}). The contours range from $k_z\phi_{u_tu_t}^+ = 0.5$ to $3.8$ with a constant increment of 1.1 units.}
\caption{Pre-multiplied two-dimensional spectra of the tangential velocity fluctuations for the NACA4412 (\textcolor{blue}{---}), NACA0012 (\textcolor{red}{---}) and ZPG (\textcolor{green}{---}) at $y_n^+ = 16$. The contours represent $k_zk_t\phi_{u_tu_t}^+ = 0.05$, $0.1$ and $0.2$.}
\caption{Pre-multiplied two-dimensional spectra of the tangential velocity fluctuations for the NACA4412 (\textcolor{blue}{---}), NACA0012 (\textcolor{red}{---}) and ZPG (\textcolor{green}{---}) at $y_n/\delta_{99} = 0.2$. The contours range from $k_zk_t\phi_{u_tu_t}^+ = 0.05$ to $0.15$ with a constant increment of 0.1 units.}
\caption{Performance comparison with deep trackers in terms of AUC. The top two values are highlighted by \textcolor{red}{red} and \textcolor{blue}{blue}.}
\caption{Attribute-based comparison on hyperspectral/color videos. The best two results are shown in \textcolor{red}{red} and \textcolor{blue}{blue} fonts. Our tracker ranks the first on 5 out of 11 attributes: IPR, OPR, DEF, BC, and OV.}
\caption{The top panel shows the luminosity distance $D_L$ vs.\redshift$z$ for $\sim 10^7$ halos within a $450$ deg$^2$ field-of-view (the color gradient illustrates the logarithmic point density) in relation to the luminosity distance $\bar{D}_L(z)$ of the homogeneous $\Lambda$CDM model at the same redshift. The bottom panel shows the bias of the source average for various distance functions.}
\caption{{\color{nicegreen} \it Left panel.} Pulls for individual observables in units of $ \sigma $. {\color{nicegreen} \it Right panel.} Status of the $ ( | V_{cb} | \, , | V_{ub} | ) $ plane. The horizontal and vertical coloured bands represent our averages of the determinations from semileptonic $ B $ decays. The white bands with solid (dashed) borders correspond to the determinations from exclusive (inclusive) semileptonic $ B $ decays. The diagonal coloured band corresponds to the determination of $ | V_{ub} |/| V_{cb} | $ from $ \Lambda_b $ decays. The rainbow oval region indicates the indirect determination of $ | V_{ub} | $ and $ | V_{cb} | $ from the global fit, without any information from semileptonic or leptonic decays of $ b $-hadrons.}
\caption{Constraint on $ | V_{td} | $ ({\color{nicegreen} \it left}), $ | V_{ts} | $ ({\color{nicegreen} \it center}), and $ | V_{td} |/| V_{ts} | $ ({\color{nicegreen} \it right}) from: the global fit (green), using only information on $ B_d $ and $ B_s $ mixings (red), and using only tree-level quantities (blue).}
\caption{ Spectroscopic redshift (\zspec ) distributions of all galaxies in the three environmental density bins considered in this work: low-density (blue), intermediate-density (green), and high-density (red). The total numbers of all galaxies in these samples are 4147, 1078, and 545 respectively. After applying the star-forming selection criteria, as described in Section \ref{sec:uvj}, the total numbers of star-forming galaxies in these samples are 3451, 729, and 229 respectively. The median \zspec\of these samples are 0.88, 0.89, and 0.84 respectively. While these redshift distributions are slightly disparate they do not correspond to significant differences in the observed redshift-evolution of the mean SFR at fixed stellar mass\citep[e.g.,][]{Speagle2014}. Nonetheless, we correct for these slight differences in redshift as described in Section \ref{sec:analysis}. }
\caption{Decay scheme of $^{50}$V. Two excited states can be populated, one via EC to \tf under emission of a 1553.77 keV \gray\and the$\beta$-decay into the first excited state of \cf resulting in a 783.29 keV \gray.}
\caption{Experimental setup for SPAD-based time-gated image sensing. The distance from the SPAD camera and laser to the target \magenta{is} 150 m. The inset photograph is \magenta{the co-registered RGB image captured and used for the non-local fusion image processing}. The series of images represent \magenta{sample preprocessed images corresponding to ten consecutive gate positions}. Depth information is gained by scanning the location of the gate.}
\caption{\textbf{Dynamic scale policy.} Scaling policies in CNNs are typically integrated into the network architecture manually in a pyramidal fashion. The color bar in this figure (second row) shows the scales at different blocks of the ResNext50 architecture. The early layers receive {\color{Blue}e\textbf{X}tra-large} resolutions and in the following layers resolutions decrease as {\color{ProcessBlue}\textbf{L}arge}, {\color{LimeGreen}\textbf{M}edium}, and {\color{ForestGreen}\textbf{S}mall}. We argue that scaling policies in CNNs should be instance-specific. Our Elastic model (the third row) allows different scaling policies for different input images and it learns from the training data how to pick the best policy. For scale challenging images e.g. images with lots of small(or diverse scale) objects, it is crucial that network can adapt its scale policy based on the input. As it can be seen in this figure, Elastic gives a better prediction for these scale challenging images. (See section~\ref{sec:policyanal} for more details)}
\caption{\small\textbf{Scale policy analysis.} This figure shows the impact of the scale policy on the accuracy of our Elastic model. (left) shows all the ImageNet validation set clustered using tsne by their scale policy pattern in the ResNeXt50+Elastic as discussed in section \ref{sec:policyanal}. (middle) shows the the scale policy score of all the images at 17 blocks of the network. Most of the images use high-resolution features at early layers and low-resolution features at later layers but some images break this pattern. Images pointed in the {\color{green}green circle} use high-resolution features in the $13^{th}$ block. Images pointed in the {\color{purple}purple circle} use low-resolution features in the $4^{th}$ block. These images usually contain a simpler pattern. (right)-bottom shows the density of images in the tsne space and (right)-top shows the density of the images that got correctly classified by Elastic model but miss-classified by the base ResNeXt model. This shows that Elastic can improve prediction when images are challenging in terms of their scale information. Some samples are pointed by the {\color{yellow}yellow circle}. Best viewed in color.}
\caption{\textcolor{red}{High-level architecture of IEC 61499 presented to implement PID control for dynamic system}}
\caption{\textcolor{red}{This explains the architecture UAC based system. The dashed arrows shows the waiting for the contract to appear. All the blocks of network are integrated in decentralized environment.}}
\caption{\textcolor{red}{This table comprehensively summarize the features, potential applications, limitations, and how bloackchain overcome those limitations of DCS and cooperative robots. It also points out the limitations of blockchain itself.}}
\caption{Material maps constructed using manual teleoperation. (Images of the environment, the ground truth and the material map created are shown for three different locations.) The color coding are as follows: \textcolor{cyan}{metal (cyan)}, \textcolor{green}{plastic (green)}, \textcolor{red}{concrete (red)}, \textcolor{blue}{cardboard (blue)}, \textcolor{yellow}{wall (yellow)}, and \textcolor{brown}{wood (brown)}.}
\caption{Construction of material map for an unknown environment: (a) Picture of the hallway used for the experiment, (b) Original map created using LiDAR from mapping robot, (c) Smoothed map , (d) Segmented boundary map, (e) Points to sample at $1m$ interval (in purple), (f) Material Map created, (g) Ground Truth material map. The color coding for material map and ground truth are as follows: \textcolor{green}{plastic (green)}, \textcolor{yellow}{ wall (yellow)}, and \textcolor{brown}{wood (brown)}.}
\caption{Rank-aware attention for skill ranking. We determine a video's rank by using high (\fgreen) and low (\fred) skill attention modules, which determine each segment's influence to the rank. Both modules are fused (\forange) for an overall skill assessment of the video. Line opacity indicates the attention value for a segment and the line thickness indicates the score.}
\caption{Rank-Aware Attention Network. Given a ranked pair of videos $(p_i, p_j)$ where $p_i$ exhibits higher skill: each video is uniformly split into segments. Extracted features (I3D) are passed into a pair of attention modules to produce video-level representations for the ranking functions (FC layers). Each ranking function produces a score $s^+$ (\fgreen) or $s^-$ (\fred). Additionally, a uniformly weighted video representation produces a third ranking score $u$ (\fblue). Three types of losses are defined: the ranking loss maximizes the margin (\fgreen-to-\fgreen, \fred-to-\fred, \fblue-to-\fblue) between the pair of ranked videos, the disparity loss ensures attention branches outperform uniform (\fgreen-to-\fblue, \fred-to-\fblue) and the final loss optimizes the attention modules to become rank-aware (\fgreen-to-\fred).}
\caption{Attention values of the high-skill (\fgreen) and low-skill (\fred) modules with the corresponding video segments for examples from `Scramble Eggs' and `Tie Tie'. The intensity of the color indicates the attention value. We show the predicted ranking from both branches.}
\caption{ Time averaged streamwise velocity profiles over longitudinal (a,b) and transversal square bars (c,d) for $N=100$ \textbf{a,c)} and $N=2.5$ \textbf{b,d)}: (\textcolor{red}{\dashed}) on the crest ($P_1$); (\textcolor{green}{\dashedb}) at a distance $0.0075h$ from the wall ($P_2$) (the cavity width is $0.05h$); (\textcolor{blue}{\dashdot}) at the center of the cavity ($P_3$); (\textcolor{black}{\solid}) velocity is also averaged in spanwise and streamwise direction. For transversal square bars, velocity profiles at the center of a lid-driven cavity from \cite{Shankar2000} are included as reference (\textcolor{magenta}{\dashdot}). }
\caption{ Dependence of the slip length (a) and slip velocity (b) on the viscosity ratio. Lines, analytical models from \cite{Schonecker2014} (solid) and \cite{Philip1972} (dashed), symbols DNS results: {\color{red} \solid}, {\color{red} \dashed}, {\color{red} $\blacksquare$} longitudinal bars; {\color{green}\solid}, {\color{green} \dashed}, {\color{green} $\bullet$} transversal square bars; {\color{magenta} $\blacktriangledown$} staggered cubes $a=0.5$, {\color{blue} $\blacktriangle$} $a=0.875$. Filled symbols are relative to two superposed fluids, empty symbols are for a single phase only. The slip length relative to arrays of cubes has been normalised by $p_c=k/(1-a)$ i.e. the equivalent pitch of 2D bars with the same fluid-area fraction of the cubes. }
\caption{ Local slip velocity above longitudinal bars (a) and transversal bars (b): (\textcolor{green}{\dashed}) free-slip boundary conditions at the interface; (\textcolor{red}{\solid}) $N=100$; ({\color{black} \chndot}) $N=2.5$; (\textcolor{blue}{\dashedb}) Stokes flow. }
\caption{Form drag ({\color{blue} \solid, $\blacktriangle$}), frictional drag on the crests plane ({\color{red} \solid}, {\color{red} $\blacksquare$}) and inside the cavities ({\color{green} \solid}, {\color{green} $\bullet$}) normalized by the total drag as a function of the viscosity ratio $N$: \textbf{a)} Longitudinal square bars, \textbf{b)} Transversal square bar and \textbf{c)} Staggered cubes. Solid lines refer to $a=0.5$, dashed to $a=0.875$. Empty symbols indicate a single fluid (plotted in correspondence of $N=1$ for analogy), solid symbols two fluids ($N$ varies).}
\caption{Dependence of the drag reduction on the viscosity ratio ($N$): {\color{red} $\blacksquare$} longitudinal bars, {\color{green} $\bullet$} transversal square bars, staggered cubes, {\color{magenta} $\blacktriangledown$ } $a=0.5$, {\color{blue} $\blacktriangle$} $a=0.875$. }
\caption{ Drag reduction as function of the slip velocity (a) and slip length in wall units (b): {\color{red} $\blacksquare$} longitudinal bars; {\color{green} $\bullet$} transversal square bars; {\color{magenta} $\blacktriangledown$} staggered cubes $a=0.5$, {\color{blue} $\blacktriangle$} $a=0.875$. Filled symbols are relative to two superposed fluids, empty symbols are for a single phase only. Solid lines are equations \ref{eq:rastegari1} and \ref{eq:rastegari2} respectively. }
\caption{ Shear stress (lines) and Reynolds stress (symbols) near the interface ($y/h=-1$) for staggered cubes (a) and longitudinal bars (b): ({\color{red} $\blacksquare$, \solid})one fluid only and ({\color{green} $\bullet$, \dashed}) LIS with $N=1$. Points are plotted every 4 for clarity. }
\caption{Root mean square of streamwise velocity (Left) $\sqrt{uu}$, and wall normal velocity (Right) $\sqrt{vv}$: \textbf{a),b)} Longitudinal Square Bars, \textbf{c),d)} Transversal Square Bars and \textbf{e),f)} Staggered Cubes $a=0.875$. Dashed lines one fluid configuration, solid lines for two superposed fluids, {\color{red} \solid} $N=1$, {\color{green} \solid} $N=2.5$. For $N=100$ both the total stress {\color{blue} \solid}, and its dispersive component {\color{black} \chndot} $\sqrt{\overline{\tilde{U_i}\tilde{U_j}}}$ are shown. A zoom near the texture is shown in the inset. { The velocity fluctuations are normalised with the bulk velocity}. }
\caption{ Maximum wall normal (a) and streamwise (b) velocity fluctuation as function of $Re_\tau$, {\color{red} $\blacksquare$} longitudinal bars; {\color{green} $\bullet$} transversal square bars; {\color{purple} $\blacktriangledown$} staggered cubes with $a=0.5$; {\color{blue} $\blacktriangle$} staggered cubes with $a=0.875$. Empty symbols indicate simulations with only one fluid without interface, solid symbols simulations with two fluids and a slippery interface, solid line Eq. \ref{eq:vvmax}, dashed line and black diamonds smooth wall data from \citep{Lee2015}.}
\caption{ Dependence of the amount of drag reduction with the maximum of the wall normal velocity rms: {\color{red} $\blacksquare$} longitudinal bars; {\color{green} $\bullet$} transversal square bars; {\color{magenta} $\blacktriangledown$} staggered cubes with $a=0.5$; {\color{blue} $\blacktriangle$} staggered cubes with $a=0.875$; solid line, Eq. \ref{eq:vvmax-fin}. Empty symbols indicate simulations with only one fluid without interface, solid symbols simulations with two fluids and a slippery interface.}
\caption{ \small Illustration of the proposed attentive RCNN. \textcolor{green}{GREEN} boxes highlight the target pair of people, and \textcolor{red}{RED} box highlights the contextual region with the highest attention. For each target pair, the attention mechanism fixates on different region. }
\caption{ \small Examples where Dual-Glance correctly predict the relationship (yellow label) while First-Glance fails (blue label). \textcolor{green}{GREEN} boxes highlight target people pair, and the top two contextual regions with highest attention are shown in \textcolor{red}{RED}. }
\caption{ \small Examples of incorrect predictions on PISC dataset. Yellow labels are the ground truth, and \textcolor{blue}{BLUE} labels are the model's predictions. }
\caption{ \small Example of correct predictions on PISC dataset. \textcolor{green}{GREEN} boxes highlight the targets, and \textcolor{red}{RED} box highlights the contextual region with highest attention. }
\caption{Network complexity and myocardial segmentation accuracy over the 5-fold cross-validation (mean and std) for the architectures studied with an input shrink factor of 2 (i.e. the original input size was reduced by a factor of 2). MSE: mean squared error. MAE: mean absolute error. BN: Batch normalization. RL: Residual learning {\color{Red} revisar...}. \label{table_accuracy}}
\caption{SUMO-generated network in UDSSC's roundabout. The blue vehicles are the AVs; they are both controlled by the RL policy. Videos of this policy in simulation are available at \textcolor{blue}{\url{https://sites.google.com/view/iccps-policy-transfer}}.}
\caption{Space-time diagrams of the simulated baseline and RL policy. Each line corresponds to a vehicle in the system. Top: a guide to the color-scheme of the space time diagrams. The northern route is in red, the western route in blue. Middle: illustrates the overlap between the merging northern platoon and the western platoon. Bottom: The RL policy, depicted at the bottom successfully removes this overlap. Videos of this policy in simulation are available at \textcolor{blue}{\url{https://sites.google.com/view/iccps-policy-transfer}}.}
\caption{\footnotesize RL-controlled vehicle demonstrating smoothing behavior in this series of images. \textbf{First}: RL vehicle slows down in anticipation of a sufficiently short inflow from the north. \textbf{Second}: The northern inflow passes through the roundabout at high velocity. \textbf{Fourth}:The RL vehicle accelerates and leads its platoon away from the roundabout. Videos of this policy in simulation are available at \textcolor{blue}{\url{https://sites.google.com/view/iccps-policy-transfer}}.}
\caption{Detailed comparisons on VOT-2016~\cite{kristan2016visual}. The best two results are highlighted in \textcolor{red}{\bf red} and \textcolor{blue}{\bf blue} fonts, respectively.}
\caption{Comparisons on VOT-2017~\cite{kristan2017visual}. The best two results are highlighted in \textcolor{red}{\bf red} and \textcolor{blue}{\bf blue} fonts, respectively.}
\caption{Comparisons on TrackingNet~\cite{muller2018trackingnet} with the best two results highlighted in \textcolor{red}{\bf red} and \textcolor{blue}{\bf blue} fonts, respectively.}
\caption{\textbf{Quantitative comparison of different model settings for saliency detection.} ``*GAP'', ``*AC'', and ``*LAP'' mean we embed these PiCANets in corresponding decoding modules. ``LC'', ``MaxP'', and ``AveP'' mean large-kernel convolution, max-pooling, and average pooling, respectively. \red{Red} indicates the best performance.}
\caption{\textbf{Quantitative evaluation of state-of-the-art salient object detection models.} \red{Red} and \blu{blue} indicate the best and the second best performance, respectively.}
\caption{\textbf{Quantitative comparison of different semantic segmentation model settings on the PASCAL VOC 2012 val set in terms of mIOU.} \red{Red} indicates the best performance in each row.}
\caption{\textbf{Quantitative comparison of different object detection model settings on the PASCAL VOC 2007 test set in terms of mAP.} ``+478LC\_910ReNet'' means we use vanilla Conv layers with large kernels for the Conv4\_3, FC7, and Conv8\_2 layers and adopt ReNet for the Conv9\_2 and Conv10\_2 layers. Other model settings can be inferred accordingly.\red{Red} indicates the best performance.}
\caption{High-Level Search of CBS{\color{red}w/P}}
\caption{(a) The noise temperature of the NW as a function of $I_\mathrm{NW}$ at $I_\mathrm{H}=0$ and $T_\mathrm{bath} = 4.2$\,K (green symbols). Solid line is the shot-noise prediction with$F = 0.32$, which is close to the universal value for elastic diffusive conductor. (b) The electron energy distribution function in the center of metallic heater, measured via the shot-noise spectroscopy, {\color{black} see Ref.~\cite{7985929} for details}. The blue symbols are extracted from measurement at $I_\mathrm{H} = 40\,\mu$A, the red symbols from measurement at $I_\mathrm{H} = 175\,\mu$A. Dashed lines are equilibrium Fermi-Dirac distribution functions with specified temperatures $T_\mathrm{H}^\mathrm{center}$. The inset shows the nonlinear differential resistance $r$ of the section between contacts 1 and 2 as a function of $I_\mathrm{H}$ at $T_\mathrm{bath}=4.2$\,K (solid orange line) and$T_\mathrm{bath}=0.5$\,K (solid blue line). The dashed line is the prediction with the conductivity correction proportional to$T^3-T_\mathrm{bath}^3$ (see text). }
\caption{Local and average noise thermometry applied to the characterization of the temperature profile in the contact heater. (a) Measured local temperature in the center of Ni/Au {\color{black} constriction} $T\mathrm{_H^{center}}$ at $T_{\rm bath}=0.5$\,K (blue diamonds) and$4.2$\,K (red diamonds) along with the corresponding numerical simulations (black lines). (b) The same as (a) but for average temperature$T\mathrm{_H^{average}}$ of the heater. The inset shows an example of simulated spatial temperature profile along the heater $T(x)$ with marked $T\mathrm{_H^{center}}$ and $T\mathrm{_H^{average}}$ {\color{black} at $I_{\rm H}=2.5$\,mA}. (c) {\color{black} $T\mathrm{_H^{center}}$ plotted as a function of $T\mathrm{_H^{average}}$. The experimental data from (a) and (b) are combined and shown by blue symbols ($T_{\rm bath}=0.5$\,K) and red symbols ($T_{\rm bath}=4.2$\,K).} The solid lines are the result of numerical calculations based on heat balance equation (see text) for both bath temperatures. {\color{black} The dashed lines for $T_{\rm bath}=0.5$\,K are predictions for the two limiting cases of the diffusion cooling and the electron-phonon cooling in a homogeneous conductor. The corresponding spatial temperature profiles$T(x)$ are shown in the nearby insets.}}
\caption{(a) The measured temperature in the center of the Ni/Au constriction $T\mathrm{_H^{center}}$ as a function of the heating current $I_\mathrm{H}$ at $T_{\rm bath}=4.2$\,K for samples with the leads composed of$n=1$ and $n=2$ layers (red symbols, see text). The dashed lines are numerical calculations with corresponding $n$, while solid line is calculated for the case of ideal leads ($n\to\infty$). The inset depicts the schematic of the {\color{black} trapezoidal leads used in the experiment}. (b) The similar measurements as (a) in a form of temperature increment $\delta T\mathrm{_H^{center}}=T\mathrm{_H^{center}}-T_{\rm bath}$ at $T_{\rm bath}=0.5$\,K (blue symbols) and$4.2$\,K (red symbols) for$n=1$ (diamonds) and $n=2$ (circles). The dashed lines are {\color{black} the fits with eq.~(\ref{Temp})} (see text). (c) The effective heater resistance $r\mathrm{_H^{*}}$ (see text) against the inverted number of layers in leads $1/n$. The theoretical predictions are built both for experimental lead geometry (solid lines) and for rectangular leads (dashed lines) at $T_{\rm bath}=0.5$\,K (blue color) and$4.2$\,K (red color). The experimental data is shown as symbols of corresponding colors (diamonds for$n=1$ and circles for $n=2$). Note, that all theoretical predictions meet at $r\mathrm{_H^{*}}=r_\mathrm{const}$ when $n\to\infty$, where $r_\mathrm{const}$ is the resistance of the {\color{black} heater constriction}.}
\caption{\label{fig3}Simulation results from jRad illustrating the key properties of radiation and degree of circular polarization from 1000 electrons trapped in a current filament. (a) and (c) show the frequency averaged spatial distribution of radiated energy on the detector lying in the $XY$ plane (parallel to the simulation plane) from the current filament shown in Fig.~\ref{fig1}c and 1b respectively. (b) and (d) show the frequency averaged degree of circular polarization corresponding to (a) and (c). For the unmagnetized filament, there is almost equal contribution to left and right circular polarization, whereas, a magnetized filament produces a strong right circularly polarized radiation with peak value $\sim 25\%$. (e) shows the frequency averaged spatial distribution of radiated energy on a detector similar to (c) from four magnetized current filaments ( with initial magnetization $\sigma$=0.05) of electron-proton plasma shown in the inset (a zoom of this inset can be found in the Appendix~Fig.~\ref{supp1}). (f) shows the frequency averaged degree of circular polarization corresponding to (e) with the inset showing the $\big<P_c\big>$ for each of the corresponding filaments.} \end{figure} To validate our model, we extracted trajectories of 1000 electrons from a current filament directly from PIC simulations and computed the radiation spectrum and the degree of circular polarization ($P_c$) of the radiation emitted from them using the post-processing radiation code jRad~\cite{jrad}. Although the OSIRIS simulation is 2D, we reconstruct the 3D trajectories as $z=\int p_z(t)/\gamma_e(t)dt$. The spectrum was calculated on a two-dimensional virtual detector in the xy plane at a distance $z=1.5\times 10^4c/\omega_{pe}$ and size $60000\times 60000(c/\omega_{pe})^2$ divided in $100\times 100$ cells to capture the entire emitting region. The detector captured a spectra of frequencies ($\omega$) in the range $\omega =(10^0-10^3)\omega_{pe}^{-1}$ with a resolution of 256 cells per decade in the frequency axis. Radiation was calculated following the trajectories from time $t_i=6300\omega_{pe}^{-1}$ to $t_f=9000\omega_{pe}^{-1}$ during which the filament was in a steady state. The degree of circular polarization ($P_c$) was estimated using the relevant Stokes parameters, in which $P_c=V/I$, $V=2\langle Im[(\boldsymbol{\epsilon}_1\cdot \textbf{E}_{rad})^*(\boldsymbol{\epsilon}_2\cdot \textbf{E}_{rad})]\rangle$, $I=|\boldsymbol{\epsilon}_1\cdot \textbf{E}_{rad}|^2+|\boldsymbol{\epsilon}_2\cdot \textbf{E}_{rad}|^2$ and $\boldsymbol{\epsilon}_1$ and $\boldsymbol{\epsilon}_2$ are the unit vectors perpendicular to the direction of observation. The angular brackets $\langle\cdot\rangle$ represent the time average. Figs.3a and 3c show the spatial distribution of frequency averaged radiated energy, $\langle I_{rad}\rangle_\omega=\int I_{rad}d\omega/\int d\omega$ from the electrons confined in the current filament of Figs.1b and 1c respectively. Figs. 3b and 3d show the corresponding $\langle P_c\rangle_\omega=\int P_cI_{rad}d\omega/\int I_{rad}d\omega$ for the filament. A strong right circular polarization with a peak value of $\langle P_c\rangle_\omega\approx 0.25 (25\%)$ is observed for the magnetized filament. For the unmagnetized filament, a nearly equal distribution of left and right circular polarization is observed resulting in a net zero circularly polarized radiation flux. The total circularly polarized radiation flux, $\langle P_c\rangle$, can be obtained using the formula $\langle P_c\rangle=\iint P_cI_{rad}dAd\omega/\iint I_{rad}dAd\omega$, where $A$ is the detector area. For the radiation in Fig.~\ref{fig1}b, $\langle P_c~\rangle=0.117 (11.7\%)$. To confirm that $\langle P_c\rangle$ was not affected by the time period on which it is averaged, the time of averaging ($6300\omega_{pe}^{-1}-9000\omega_{pe}^{-1}$) was divided into three equal segments of interval $900\omega_{pe}^{-1}$ and $\langle P_c\rangle$ was separately calculated for each of these intervals. We found that $\langle P_c\rangle$ was equal in all the cases with a variation of $\pm 0.1\%$, which indicates that our observation for a single filament are physically meaningful. Furthermore, to check for the effect of multiple filaments, we computed the radiation spectra and $P_c$ from 1000 electrons distributed equally in four filaments of a magnetized electron-proton plasma of Fig.~\ref{fig1}b. Figs.~\ref{fig3}e and ~\ref{fig3}f show the $\langle I_{rad}\rangle_\omega$ and $\langle P_c\rangle_\omega$ respectively for the four filaments described in the inset of Fig.~\ref{fig3}e. The radiation spectra and circular polarization is similar to the single filament. In addition, the $\langle P_c\rangle$ obtained from each of the filaments (inset of Fig.~\ref{fig3}f) is computed separately and found to be nearly equal ($\sim 0.115$). This shows that the anisotropy in pitch angle distribution depends only on the fields and not the filament shape. An important observation is that WI/CFI in magnetized plasmas generate local islands of anisotropic pitch angle distribution. Hence, circular polarization can be observed even in case of WI/CFI driven shocks in magnetized spherical jets irrespective of the position of observation in contrast to the model described earlier~\cite{nava}, which is valid only for collimated jets. \begin{figure} \centering \noindent\includegraphics[width=3.0in]{fig4.pdf} \caption{\label{fig4}Simulation results from jRad illustrating the normalized flux of circularly polarized photons (averaged over frequency and spatial domain) from 1000 electrons trapped in a typical current filament arising from interpenetrating flows of electron-positron and electron-ion plasmas for varying magnetizations. Note that here negative $\sigma$ represents negative $B_z$.} \end{figure} To investigate the role of initial magnetization of the plasma, $\langle P_c\rangle$ was calculated from 1000 electrons confined in a single filament of electron-proton plasma with initial magnetizations varying from 0.0 to 0.2 for both signs of $B_z$. It was found that the $\langle P_c\rangle$ increased initially with magnetization and then started to saturate beyond $\sigma = 0.05$ converging to $\sim 13\%$. % The level of circular polarization depends on the electron pitch angle distribution, and we have confirmed that the external magnetic field changes this distribution such that finite levels of circular polarization become possible. Specifically, in unmagnetized scenarios, the pitch angle distribution is symmetrical about $\pi/2$. Thus, the total level of circular polarization vanishes in this case. In magnetized scenarios, instead, the pitch angle distribution is off-set due to the particle drifts that appear in the presence of the external magnetic fields. As the pitch angle distribution becomes asymmetric, the level of circular polarization level increases. %% Other processes can also lead to finite circular polarization levels, for instance \cite{sinha} where generation of circular polarization in a different physical configuration is attributed to an assymetric energy dissipation mechanism instead of the topological changes in pitch-angle distribution discussed here. %This indicates that the electron velocity vector in the azimuthal direction competes with the velocity vector due to the cyclotron motion induced by the initial magnetization. The influence of cyclotron motion due to initial magnetizaton on the pitch angle increases with $\sigma$ leading to an increase in the degree of circular polarization. For high magnetizations ($\gtrapprox 0.05$), the cyclotron motion completely dominates over the azimuthal velocity vector leading to a saturation of the pitch angle distribution which results in the saturation of the $\langle P_c\rangle$. The handedness of circular polarization changes with the sign of $B_z$ because the velocity vector due to cyclotron motion reverses with the sign of $B_z$ resulting in an anisotropy in the opposite direction. Furthermore, to understand the effect of plasma composition i.e. ion-electron mass ratios ($m_i/m_e$), we performed simulations for the same magnetizations for plasmas with $m_i/m_e=1$ and 918. We observed that the values of $\langle P_c\rangle$ for $m_i/m_e=918$ and 1836 were almost equal. This is because the ion response time scales are very large when compared to electron time scales, resulting in the radiation primarily emitted only by the electrons and the circular polarization arising only due to the anisotropies in electron pitch angle distribution. For electron-positron plasmas, both species contribute to the emitted radiation. They rotate in opposite directions due to their opposite charge and create anisotropies in mutually opposite directions resulting in a net cancellation of the circular polarization. In conclusion, we have shown that the motion of the charged particles in the fields due to WI/CFI produces strong radiation emission in the direction of the plasma flow. An initial magnetization breaks the symmetry in pitch angle distribution of the electrons resulting in a partially circularly polarized radiation. Results indicate that the anisotropies in pitch angle distribution saturates at high initial magnetizations leading to a saturation in the degree of circular polarization. The emission of circularly polarized radiation is limited to electron-ion plasmas (plasmas with $m_i/m_e > 1$). For pair plasmas, the electrons and positrons produce circular polarization of opposite handedness resulting in a net zero circular polarization. The simulation set-up mimics the flow believed to be present in astrophysical scenarios. Hence, this study is significant to understand the origin of circular polarization in the recent observation of GRB afterglow~\cite{wiersema2014}. As the anisotropy in pitch angles is local to the current filaments, the treatment is valid even for broader or spherical jets. With the observation of WI/CFI in laboratory and future proposals to study the interaction of electrically neutral electron-positron beams (fireball beams) with plasma~\cite{huntington,fox2013,sarri}, it may be possible to design scaled experiments to test the results presented in this Letter and develop a better understanding of radiation and its polarization in real astrophysical scenarios. %We finally note that there are other configurations that allow for the production of circularly polarized radiation in WI/CFI scenarios based on a different physical mechanism (Sinha et al. to be submitted), thus stressing that the WI/CFI may play an important role in explaining the origin of circularly polarized radiation recently observed~\cite{wiersema2014}. %% If you wish to include an acknowledgments section in your paper, %% separate it off from the body of the text using the \acknowledgments %% command. \acknowledgments The authors would like to thank Prof. Mikhail Medvedev for very insightful discussions. This work was supported by the European Research Council (ERC-2010-AdG grant 267841 and ERC-2015 grant 695008) and FCT (Portugal) grant no. PTDC-FIS-PLA-2940-2014. J.V. acknowledges the support of FCT (Portugal) grant no. SFRH/IF/01635/2015. We acknowledge PRACE for awarding access to resource SuperMUC (Leibniz research center) and Fermi (CINECA). We also acknowledge the supercomputing resource IST Cluster at IST. \software{OSIRIS \citep{osiris,fonseca2013}, jRad \citep{jrad}} %% To help institutions obtain information on the effectiveness of their %% telescopes the AAS Journals has created a group of keywords for telescope %% facilities. % %% Following the acknowledgments section, use the following syntax and the %% \facility{} or \facilities{} macros to list the keywords of facilities used %% in the research for the paper. Each keyword is check against the master %% list during copy editing. Individual instruments can be provided in %% parentheses, after the keyword, but they are not verified. \vspace{5mm} %\facilities{HST(STIS), Swift(XRT and UVOT), AAVSO, CTIO:1.3m, %CTIO:1.5m,CXO} %% Similar to \facility{}, there is the optional \software command to allow %% authors a place to specify which programs were used during the creation of %% the manusscript. Authors should list each code and include either a %% citation or url to the code inside ()s when available. %\software{astropy \citepp{2013A&A...558A..33A}, % Cloudy \citepp{2013RMxAA..49..137F}, % SExtractor \citepp{1996A&AS..117..393B} % } %% Appendix material should be preceded with a single \appendix command. %% There should be a \section command for each appendix. Mark appendix %% subsections with the same markup you use in the main body of the paper. %% Each Appendix (indicated with \section) will be lettered A, B, C, etc. %% The equation counter will reset when it encounters the \appendix %% command and will number appendix equations (A1), (A2), etc. The %% Figure and Table counter will not reset. \appendix %\counterwithin{figure}{section} \section*{Degree of circular polarization from multiple filaments} %\subsection{} Although we have only considered radiation from a single filament, it is important to consider the net effect on the degree of circular polarization due to radiation from multiple filaments. A simple analytic model can be constructed by assuming the current filaments as independent sources of radiation. In the far field, each filament can be described as a point source of radiation with electric field components given by $E_x(r,t)=A\exp[i(\textbf{k}\cdot\textbf{r}-\omega t)]$ and $E_y(r,t)=A\exp[i(\textbf{k}\cdot\textbf{r}-\omega t+\phi)]$ with $\textbf{k}$ being the wave vector, $\omega$ the frequency and $\phi$ the phase difference between the components. The degree of circular polarization associated with a single filament is then, $P_c=\sin(\phi)$ where the light is circularly (elliptically) polarized for $\phi=\pi/2(\phi<\pi/2)$. To incorporate the effect of multiple current filaments, we consider the superposition of $N$ plane waves with a random phase factor $\psi_k$ between them. The resultant electric field components can be written as $E_x(r,t)=\sum_{n=1}^{N}A_{n}\exp[i(\textbf{k}_{n}\cdot\textbf{r}-\omega_{n}t+\psi_{n})]$ and $E_y(r,t)=\sum_{m=1}^{N}A_{m}\exp[i(\textbf{k}_m\cdot\textbf{r}-\omega_{m}t+\psi_{m}+\phi_m)]$. The degree of circular polarization ($P_c$) is given by the relevant Stokes parameters, in which $P_c=V/I$, $V=2\langle \mathrm{Im}\{E_x^*E_y\} \rangle$, $I=E_x^* E_x + E_y^* E_y$, and where the angular brackets $\langle\cdot\rangle$ represent the time average. Hence, $V = 2 \langle \textrm{Im} \{ \sum_{n,m=1}^{N} A_n A_m \exp \left[i \left(\Delta \mathbf{k}_{n,m}\cdot\textbf{r}-\Delta \omega_{n,m} t+\Delta \psi_{n,m}+\phi_m\right)\right]\}
\caption{\protect Comparison of EDSR* with EDSR~\cite{lim2017enhanced} on benchmark datasets. {\color[HTML]{FE0000}Red} indicates better results than the other.}
\caption{\protect Comparisons of state-of-the-art methods on speed with DIV2K dataset. {\color[HTML]{FE0000}Red} indicates the best results and the {\color[HTML]{3531FF}blue} indicates the second best results.}
\caption{\protect Comparison of the SRGAN and Ours on $\times$4 and $\times$8 enlargement. {\color[HTML]{FE0000}Red} indicates the best results.}
\caption{\small 3D object detection results on the \textbf{pedestrian} and \textbf{cyclist} categories on the \emph{test} set. We compare pseudo-LiDAR with \PSMNetpd (in {\color{blue}blue}) and LiDAR (in {\color{gray} gray}). We report \APBEV~/ \AP at IoU = 0.5 (the standard metric). $\dagger$: Results on the KITTI leaderboard.}
\caption{\textbf{Pseudo-LiDAR signal from visual depth estimation.} Top-left: a KITTI street scene with super-imposed bounding boxes around cars obtained with LiDAR ({\color{red}red}) and pseudo-LiDAR ({\color{green}green}). Bottom-left: estimated disparity map. Right: pseudo-LiDAR ({\color{blue}blue}) vs. LiDAR ({\color{yellow}yellow}) --- the pseudo-LiDAR points align remarkably well with the LiDAR ones. Best viewed in color (zoom in for details).}
\caption{\textbf{\PSMNet vs. \PSMNetpd.} Top: a KITTI street scene. Left column: the depth map and pseudo-LiDAR points (from the bird's-eye view) by \PSMNet, together with a zoomed-in region. Right column: the corresponding results by \PSMNetpd. The observer is on the very right side looking to the left. The pseudo-LiDAR points are in {\color{blue}blue}; LiDAR points are in {\color{yellow}yellow}. The pseudo-LiDAR points by \PSMNet have larger deviation at far-away distances. Best viewed in color (zoom in for details).}
\caption{\textbf{Qualitative comparison and failure cases.} We compare \AVOD with LiDAR, pseudo-LiDAR (stereo), pseudo-LiDAR (monocular), and frontal-view (stereo). Ground-truth boxes are in {\color{red}{red}}; predicted boxes in {\color{green}{green}}. The observer in the pseudo-LiDAR plots (bottom row) is on the very left side looking to the right. The \textbf{mislocalization} cases are indicated by {\color{gray}gray} arrows; the \textbf{missed detection} cases are indicated by {\color{yellow}yellow} arrows. The frontal-view approach (\emph{bottom-right}) makes extremely inaccurate predictions, even for nearby objects. Best viewed in color.}
\caption{\textbf{Qualitative comparison and failure cases.} We compare \AVOD with LiDAR and pseudo-LiDAR (stereo). Ground-truth boxes are in {\color{red}{red}}; predicted boxes in {\color{green}{green}}. The observer in the pseudo-LiDAR plots (bottom row) is on the bottom side looking to the top. The pseudo-LiDAR-based detection misses the partially occluded car (the {\color{yellow} yellow} arrow), which is a hard case. Best viewed in color.}
\caption{\textbf{Pseudo-LiDAR signal from visual depth estimation.} Top-left: a KITTI street scene with super-imposed bounding boxes around cars obtained with LiDAR ({\color{red}red}) and pseudo-LiDAR ({\color{green}green}). Bottom-left: estimated disparity map. Right: pseudo-LiDAR ({\color{blue}blue}) vs. LiDAR ({\color{yellow}yellow}) --- the pseudo-LiDAR points align remarkably well with the LiDAR ones. Best viewed in color (zoom in for details.)}
\caption{\small 3D object detection results on the KITTI validation set. We report \APBEV ~/ \AP (in \%) of the \textbf{car} category, corresponding to average precision of the bird's-eye view and 3D object box detection. Mono stands for monocular. Our methods with \emph{pseudo-LiDAR} estimated by \PSMNetpd~\cite{chang2018pyramid} (stereo) or \DORN~\cite{fu2018deep} (monocular) are in {\color{blue} blue}. Methods with LiDAR are in {\color{gray} gray}. Best viewed in color.}
\caption{\small 3D object detection on the \textbf{pedestrian} and \textbf{cyclist} categories on the validation set. We report \APBEV~/ \AP at IoU = 0.5 (the standard metric) and compare \Frustum with \emph{pseudo-LiDAR} estimated by \PSMNetpd (in {\color{blue}blue}) and LiDAR (in {\color{gray} gray}). }
\caption{\textbf{Qualitative comparison.} We compare \AVOD with LiDAR, pseudo-LiDAR, and frontal-view (stereo). Ground-truth boxes are in {\color{red}{red}}, predicted boxes in {\color{green}{green}}; the observer in the pseudo-LiDAR plots (bottom row) is on the very left side looking to the right. The frontal-view approach (\emph{right}) even miscalculates the depths of nearby objects and misses far-away objects entirely. Best viewed in color.}
\caption{\small 3D object detection results on the \textbf{car} category on the \emph{test} set. We compare \emph{pseudo-LiDAR} with \PSMNetpd (in {\color{blue}blue}) and LiDAR (in {\color{gray} gray}). We report \APBEV~/ \AP at IoU = 0.7. $\dagger$: Results on the KITTI leaderboard.}
\caption{\small Overview of our SymmFCNet. {\color{red}Red}, {\color{green}green} and {\color{blue}blue} lines represent the pixel-wise correspondence between the input and the flip image. {\color{red}Red:} missing pixels (input) to non-occluded pixels (flip); {\color{green}Green:} missing pixels (input) to missing pixels (flip); {\color{blue}Blue:} remaining pixels (input) to remaining pixels (flip). }
\caption{\textcolor{blue}{Best} completion results with RankGAN on CelebA, `Center Large' mask.}
\caption{\textcolor{blue}{Best} completion results with RankGAN on CelebA, `Center Large' mask.}
\caption{\textcolor{red}{Worst} completion results with RankGAN on CelebA, `Center Large' mask.}
\caption{\textcolor{blue}{Best} completion results with RankGAN on CelebA, `Center Small' mask.}
\caption{\textcolor{red}{Worst} completion results with RankGAN on CelebA, `Center Small' mask.}
\caption{\textcolor{blue}{Best} completion results with RankGAN on CelebA, `Periocular Large' mask.}
\caption{\textcolor{red}{Worst} completion results with RankGAN on CelebA, `Periocular Large' mask.}
\caption{\textcolor{blue}{Best} completion results with RankGAN on CelebA, `Periocular Small' mask.}
\caption{\textcolor{red}{Worst} completion results with RankGAN on CelebA, `Periocular Small' mask.}
\caption{Input is \textcolor{blue}{positive} correlated Gaussian}
\caption{Input is \textcolor{blue}{general} correlated Gaussian}
\caption[]{\label{Figure2}\bf Hybridized mode. a\sf, Scanning electron micrograph of a resonantly driven ($f_\text{LV}=\SI{7.527}{\mega\hertz}$) pair of nanopillars ($H\approx\SI{7}{\micro\metre}$, $r\approx\SI{310}{\nano\metre}$, $d\approx\SI{1.3}{\micro\metre}$) imaged from the top and \bf b, \sf in a $\SI{60}{\degree}$ tilted view. Red and blue dotted lines indicate the circumference of the undriven left and right resonator, respectively. \bf c\sf, Amplitude for different drive frequencies $f_{\text{drive}}\approx f_{\text{LV}}$ of left (\tikz\node[draw,circle,very thick,fill=red, inner sep=0pt,minimum size=7pt]{};) and right (\tikz\node[draw,very thick,fill=blue, inner sep=0pt,minimum size=7pt]{};) resonator determined from the scanning electron micrographs. Scale bar in \bf a \sf and \bf b \sf corresponds to $\SI{1}{\micro\metre}$. }
\caption[]{\label{Figure4}\bf Geometry dependence of coupling strength. a\sf, Experimentally determined coupling rate $g/2\pi$ of the vertical modes of the two nanopillars over their center-to-center distance $d$ for several samples and pillar pair geometries with $r\approx\SI{430}{\nano\metre}$ \&$H\approx\SI{7}{\micro\metre}$ (red), $r\approx\SI{335}{\nano\metre}$ \&$H\approx\SI{7}{\micro\metre}$ (blue), $r\approx\SI{330}{\nano\metre}$ \&$H\approx\SI{8.2}{\micro\metre}$ (green). Error margins describe fitting errors. The inset shows the vibration direction for differently oriented pillar pairs (see Supplementary Information for details). \bf b, c \sf and \bf d\sf, Finite element simulation of level splitting depending on \bf b \sf center-to-center distance $d$ ($r=\SI{400}{\nano\metre}$, $H=\SI{7}{\micro\metre}$), \bf c \sf nanopillar bottom radius $r$ ($d=2r+\SI{400}{\micro\metre}$, $H=\SI{7}{\micro\metre}$) and \bf d \sf nanopillar height $H$ ($r=\SI{400}{\nano\metre}$, $d=\SI{1.2}{\micro\metre}$). We assume an isotropic substrate with Young's Modulus $E_{\text{[100]}}=\SI{85.9}{\giga\pascal}$, and we display vertical (\tikz\node[circle ,fill=black, inner sep=0pt,minimum size=8pt]{};) and horizontal (\tikz\node[draw,circle,very thick,fill=none, inner sep=0pt,minimum size=7pt]{};) modes. }
\caption{A comparison between \textit{HC} and the spectrogram of the cry signal. The horizontal line is the corresponding noise threshold $\overline{r}$. The green intervals are classified as cry units in the first segmentation and the vertical bars {\color{blue}| }represent the respective cut-points.\label{cshvsspec}}
\caption{Spectrogram of the first significant cry unit, \textit{FC}. Each {\color{red}$\ast$} represents the fundamental frequency in the respective window estimated by the corresponding high-frequency ripple in the cepstrogram of \textit{FC}. The green line is the mean fundamental frequency of the first cry \textit{F0}.\label{FC}}
\caption{ (a) Decay form factors for different channels Symbols represent the source sink separations : $t_{sep} = 8 $ (\protect\markerdiamond), 9(\protect\markercircle), and 10(\protect\invtriangle). (b) Comparison of normalized form factors with our data at $t_{sep} = 8$ and the earlier study \cite{Aoki:2017puj}. }
\caption{\color{Gray} \textbf{A linear dynamical system can be used to accurately estimate the average stochastic spiking of neurons in a branching process model.} \textbf{a,} An example schematic of the cascade generated by an external stimulus shown in the leftmost portion of panel \textbf{b}. The red nodes indicate active nodes, and the gray nodes indicate nodes that have already been active. \textbf{b,} Examples of different cascades generated by stimulating the same node of a fixed weighted random network. In this case, we stimulate node 8 and note that therefore node 8 is the only active node in the first time bin, but thereafter the cascades may differ. \textbf{c,} The activity of each node and time step, averaged over 1,000 simulations on the same weighted random network and with the same stimulation node used to produce the results shown in panel \textbf{b}. \textbf{d,} The linear dynamical system with the same weighted random network produces an estimation for spike counts of the stochastic branching model. \textbf{e,} A histogram of the difference between the cascades produced by the simulation average shown in panel \textbf{c} and the spike counts stipulated by the linear dynamical system shown in panel \textbf{d}. The normal fit has a mean of -6.4 $\times$ 10\textsuperscript{-4} and a standard deviation of 4.2 $\times$ 10\textsuperscript{-3}. \textbf{f,} For weighted random networks of size 50, 100, 150, 200, 250, and 300 nodes, all with fractional connectivity of 0.2, we show the differences between the average cascade activity and the estimation via linear dynamics. The error bars indicate standard deviations, and the means are $2.094 \times 10^{-4}$ (50 nodes), $-6.76 \times 10^{-5}$ (100 nodes), $-9.7 \times 10^{-6}$ (150 nodes), $-8.0 \times 10^{-5}$, $5.75 \times 10^{-5}$ (200 nodes), $5.75 \times 10^{-5}$ (250 nodes), and $-2.05 \times 10^{-5}$ (300 nodes). }
\caption{\color{Gray} \textbf{Network topology constrains cascade duration.} \textbf{a,} The network activity of an example cascade on a weighted random network. The cascade duration is defined as the number of time steps $t$ between the point at which the first spike occurs after a time step of quiescence, and the point at which the last spike occurs, followed by a time step of quiescence. \textbf{b,} The dynamics of the stochastic model can be accurately predicted by linear estimation. The plot shows the fraction of cascades alive at time $t$ from 1,000 trials of stimulating a node in a single instantiation of a 10-node weighted random graph. The root-mean-square error between the stochastic simulation and the linear prediction is 0.0036. \textbf{c,} Networks with larger dominant eigenvalues allow longer cascades. The plot shows the mean duration of cascades as a function of the dominant eigenvalue of the network. Each of the 100 points reflects a single weighted random network composed of 10 nodes. The value of each point is the mean across the 1,000 values obtained from stimulating each of 1,000 randomly selected nodes from the graph. Note that because the network is composed of 10 nodes, and 1,000 selections are made, each node is selected multiple times. \textbf{d,} Networks with larger sum of eigenvalues allow longer cascades. The plot shows the mean duration of cascades as a function of the sum of eigenvalues of the network. Similar to panel \textbf{c}, each of the 100 points reflects a single weighted random network composed of 10 nodes. The value of each point is the mean across the 1,000 values obtained from stimulating each of 1,000 randomly selected nodes from the graph. }
\caption{\color{Gray} \textbf{Cycles and strong connections support avalanching dynamics.} \textbf{a,} The probability distribution of cascade duration in an acyclic network where all edges have a weight of 0.5. In $10^4$ trials, the maximum cascade duration is 3 time steps. \textbf{b,} The probability distribution of cascade duration in a cyclic network where all edges have a weight of 0.5. In $10^4$ trials, the maximum cascade duration is 13 time steps. \textbf{c,} Networks with higher cycle density have longer cascades. The cycle density is the number of simple cycles divided by the number of edges. A directed acyclic graph was randomly rewired to produce networks of varying cycle density. The Pearson's correlation coefficient is $r=0.8180$, $p = 1.0835\times 10^{-27}$. \textbf{d,} A schematic of a 2-node network. The weights are redistributed from the 2-cycle to self-loops by $\Delta w$. \textbf{e,} Three probability distributions of cascade duration for $\Delta w = 0.02$, $0.24$, and $0.46$ in the 2-node cycle. The distributions have slopes of $-2.2551$, $-5.6259$, and $-6.4050$, respectively. \textbf{f,} The slope of distributions (\textit{left}) decreases as $\Delta w \rightarrow 0.5$ and (\textit{right}) increases again as $\Delta w \rightarrow 1$. The quadratic fit of the slope of duration is $y = 9.9509\Delta w^2 -10.1443\Delta w -2.9108$. \textbf{g,} Network geometries producing longer duration can be characterized by a larger sum of eigenvalues; note that these networks have the same dominant eigenvalues. The Pearson's correlation coefficient is $r=0.6926$, $p=3.5190\times 10^{-8}$. The y-axis is the slope of duration as in panel \textbf{f}. \textbf{h,} A schematic of a 4-node network. The weights are redistributed from the 4-cycle to randomly chosen edges that are not part of the original 4-cycle by $\Delta w$. \textbf{i,} Three probability distributions of cascade duration for $\Delta w =$ 0.02, 0.24, and 0.46 in the 4-node cycle. The distributions have slopes of -1.0870, -2.6681, and -4.1164, respectively. \textbf{j,} The slopes of distributions follow the same shape as in panel \textbf{f}. The quadratic fit of the slope of duration is $y = 6.7078\Delta w^2 -9.3683\Delta w -1.0733$. Note: The schematic (panel \textbf{h}) and simulations (panels \textbf{i-k}) use one random selection of edges that are not part of the original cycle. \textbf{k,} The 4-node cycle shows the same relationship between duration and sum of eigenvalues as in the 2-node cycle show in panel \textbf{g}. The Pearson's correlation coefficient is $r=0.7611$, $p=2.1791 \times 10^{-10}$. The y-axis is the slope as in panel \textbf{j}. }
\caption{\color{Gray} \textbf{Network controllability is tightly linked with dynamics.} \textbf{a-c,} Scatter plots between mean cascade duration and \textbf{a} the magnitude of the eigenprojection, \textbf{b} the modal controllability, and \textbf{c} the finite average controllability. The network used here has 100 nodes with weighted random graph topology and a fractional connectivity of around 0.2. Each point is the mean duration and metric for each node. \textbf{d,} Pearson's correlation coefficients for 30 random instantiations of networks with the same parameters as the network used in panels \textbf{a-c}. \textbf{e-g,} Scatter plots between mean cascade duration and \textbf{e} the magnitude of the eigenprojection, \textbf{f} the modal controllability, and \textbf{g} the finite average controllability. The network used here has the same parameters as networks from panels \textbf{a-c}, but with 10\% of connections normally distributed with a mean of 0.9 and 90\% of connections with a mean of 0.1, all with a standard deviation of 0.1, before weight normalization. \textbf{h,} Pearson's correlation coefficients for 30 random instantiations of networks with the same parameters as the network used in panels \textbf{e-g}. }
\caption{\color{Gray} \textbf{A stimulus can be well-recovered when it generates long-lasting cascades.} \textbf{a-b,} A schematic showing two cascades triggered by different stimuli. \textbf{c,} Recovery of the stimulus using an observation of a network state during a cascade. \textbf{d,} Failed recovery of the stimulus using an observation of a network state during a cascade. \textbf{e-h,} Decay in mutual information (MI) over time. When activity from a stimulus pattern lasts longer, mutual information also persists for longer. Panels \textbf{e-h} show results from four graph types: a weighted random graph, a random geometric graph, a modular graph with 4 communities, and a Watts-Strogatz graph. We used a fractional connectivity of 0.05 to show a wide range of decay rates in mutual information (see Supplemental Materials for simulations with other fractional connectivities). \textbf{i,} The linear decay rate of mutual information over the first 10 time steps plotted against the mean cascade duration in the example weighted random graph from panel \textbf{e}. \textbf{j,} A boxplot of the Pearson correlation coefficients between the linear slope of decay in mutual information over time and the mean cascade duration. The boxplot shows data from 30 instantiations of each graph type, each network containing 100 nodes and characterized by a fractional connectivity of around 0.05. \textbf{k,} The mean decay rate in mutual information for a network is correlated with the sum of eigenvalues of the network (Pearson's correlation coefficient $r=0.9188$, $p=1.8 \times 10^{-49}$). }
\caption{CNN hyperparameter options \label{alg:Suppl_List-of-CNN-hyperparameter-options} (see Figure \ref{fig:Suppl_CNN-generic-model/architecture}). Best parameters are colored in \textcolor{blue}{blue}.}
\caption{CNNWide hyperparameter options \label{alg:Suppl_List-of-CNNWide-hyperparameter-options} (see Figure \ref{fig:Suppl_CNNWide-generic-model/architecture}). Best parameters are colored in \textcolor{blue}{blue}.}
\caption{NN hyperparameter options \label{alg:Suppl_List-of-NN-hyperparameter-options} (see Figure \ref{fig:Suppl_NN-generic-model/architecture}). Best parameters are colored in \textcolor{blue}{blue}.}
\caption{% Neural persistence values of trained perceptrons~(\textcolor{printable_1}{green}), diverging ones~(\textcolor{printable_2}{yellow}), random Gaussian matrices~(\textcolor{printable_3}{red}), and random uniform matrices~(\textcolor{printable_4}{black}). % We performed 100 runs per category; % dots indicate neural persistence while crosses indicate the predicted lower bound according to Theorem~\ref{thm:Tighter bounds}. % The bounds according to Theorem~\ref{thm:Neural persistence bounds} are shown as dashed lines. }
\caption{% Comparison of mean normalized neural persistence for trained networks without modifications~(\textcolor{printable_1}{green}), with batch normalization~(\textcolor{printable_2}{yellow}), and with 50\% of the neurons dropped out during training~(\textcolor{printable_3}{red}) for the `MNIST' data set~(50 runs per setting). }
\caption{% Traditional graph measures~(top), such as the clustering coefficient, fail to detect differences in the complexity of neural networks. % Our novel \emph{neural persistence} measure~(bottom), by contrast, shows that trained networks with $\eta = 0.5$~(\textcolor{printable_1}{green}), which have an accuracy of $\approx 0.91$, obey a different distribution than networks trained with $\eta = \num{1e-0.5}$~(\textcolor{printable_2}{yellow}), which have accuracies ranging from $0.38$--$0.65$. }
\caption{% Comparison of test set accuracy for trained networks without modifications~(\textcolor{printable_1}{green}), with batch normalization~(\textcolor{printable_2}{yellow}), and with 50\% of the neurons dropped out during training~(\textcolor{printable_3}{red}) for the MNIST data set. }
\caption{The left and right panels show direct photon $p_{T}$ spectra in p+p and in the central 0-5\% p+Au collisions at 200\,GeV (external conversions with VTX). The PHENIX previously measured p+p and minimum bias d+Au cross sections from an internal conversion method\cite{Adare:2013} are included in this figure and shown in the left panel (with p+p), and in the central panel (with p+p and d+Au).}
\caption{ Fs-laser fabrication setup which used \blue{Schwarzschild} reflection objective lens of numerical aperture $NA = 0.5$.}
\caption{(a) A phase characterisation setup of the optical vortex beams generated by $q$-plates. (b) Experimental interferograms of different $q$-plates. \blue{Scale bars 1~mm.} (c) Numerical simulations of the phase of the corresponding LG vortex beams (Eqn.~\ref{e3}). }
\caption{\label{fig:RMSE-quali} Values simulated for the 104 models ({\color{green} green lines ---}), observations ({\color{red} red solid line~---}), one-step-ahead forecasts by Ridge ({\color{blue} blue solid line~---}) and EWA (black dotted line - - -).}
\caption{\label{fig:RMSE-qualiD} Values simulated for the 104 models ({\color{green} green lines ---}), observations ({\color{red} red solid line~---}), one-step-ahead forecasts by Ridge ({\color{blue} blue solid line~---}) and EWA (black dotted line - - -).}
\caption{\label{fig:RMSE-quanti} RMSEs of the best model ({\color{strongyellow} yellow bars}) and of the best convex combination of models ({\color{pink} pink bars}) for each property, as well as the RMSEs of the considered algorithms: Ridge (top graphs) and EWA (bottom graphs). The RMSEs of the latter are depicted in {\color{blue} blue} whenever they are smaller than that of the best model for the considered time-series, in {\color{red} red} otherwise.}
\caption{\label{fig:Lasso-P1} Values simulated for the 104 models ({\color{green} green lines ---}), observations ({\color{red} red solid line~---}) and one-step-ahead forecasts by Lasso ({\color{purp} purple solid line ---}). }
\caption{\label{fig:Lasso-P2} RMSEs of the best model ({\color{strongyellow} yellow bars}) and of the best convex combination of models ({\color{pink} pink bars}) for each property, as well as the RMSEs of Lasso. The latter are depicted in {\color{blue} blue} whenever they are smaller than that of the best model for the considered property, in {\color{red} red} otherwise. }
\caption{The model for deconfounding reinforcement learning. Grey nodes denote observed variables and white nodes represent unobserved variables. \textcolor{red}{Red} and \textcolor{blue}{blue} arrows emphasize the observed variables affected by $u$ and by $z_t$, respectively. The causal effects of interest are colored in \textcolor{green}{green}.}
\caption{\label{fig:gen_test_examples}\label{fig:gen_examples}Qualitative results on the generalization test set created from ScanNet. The supervised approach is able to predict correctly large parts of the segmentations on new scenes but still makes many mistakes. S4-Net can exploit additional non-labeled sequences of other scenes to correct most of these mistakes. The floor class is shown in {\color{floor}\textbf{green}}, the structure class in {\color{structure} \textbf{blue}}, the furniture class in {\color{furniture}\textbf{cyan}}, and the props class in {\color{props}\textbf{yellow}}. }
\caption{(a) Plot of the electric field as a function of time for the Sauter pulse~\protect\eqref{Sauter pulse} as external field for $E_0=1E_c$, $\tau=1/m$ and $\alpha=0.05$ (\textcolor{blue}{blue}), $\alpha=0.1$ (\textcolor{cyan}{cyan}), $\alpha=0.2$ (\textcolor{red}{red}). The inset is a zoom on the peak around $t=0$, where additionally the external electric field is plotted by the black dashed curve. (b) The total pair number for $\alpha=0.2$ (\textcolor{orange}{orange}) and $\alpha=0$, i.e.\without backreaction (black dashed). (c) The same as in (a) but only for$\alpha=0.2$ (\textcolor{red}{red}) over a large time interval.}
\caption{(a) Plot of the electric field as a function of time for the $C^\infty$ pulse~\protect\eqref{const bump pulse} as external field for $E_0=0.1E_c$, $t_r=2/m$, $\alpha=0.1$ and $t_f=10/m$ (\textcolor{blue}{blue}), $t_f=20/m$ (\textcolor{cyan}{cyan}), $t_f=50/m$ (\textcolor{red}{red}). The external electric field is plotted in black dashed for $t_f=50/m$. (b) The total pair number for $t_f=50/m$ (\textcolor{orange}{orange}) and without backreaction (black dashed1). (c) The same as in (a) but only for $t_f=50/m$ (\textcolor{red}{red}) over a large time interval.}
\caption{Plot of the electric field as a function of time for the $C^\infty$ pulse~\protect\eqref{const bump pulse} as external field for $E_0=0.1E_c$, $t_r=2/m$, $t_f=200/m$ and $\alpha=0.1$ (\textcolor{red}{red}), $\alpha=0.2$ (\textcolor{cyan}{cyan}), $\alpha=0.5$ (\textcolor{blue}{blue}). The external electric field is plotted in black dashed.}
\caption{General statistics on each scene, showing lower average errors and deviation in the case of localisation using pose estimation. {\color{red} add \% missing}}
\caption{Comparison with state-of-the-art methods on Market1501, CUHK03 (new protocol, \textit{detected} subset) and DukeMTMC-reID, under single-domain setting. Methods with * require keypoint and (or) segmentation assistance. In each column, the \textcolor{red}{1st} and \textcolor{blue}{2nd} highest scores (excluding methods with trailing \dag) are marked by \textcolor{red}{red} and \textcolor{blue}{blue}, respectively.}
\caption{Comparison with state-of-the-art methods under cross-domain setting. In each column, the \textcolor{red}{1st} and \textcolor{blue}{2nd} highest scores (excluding methods with trailing \dag) are marked by \textcolor{red}{red} and \textcolor{blue}{blue}, respectively. \textbf{DT} means Direct Transfer.}
\caption{Various wave and interface locations during a Mach 1.22 shock-R22 cylindrical bubble interaction in addition to a small schematic of this interaction. The figure compares experimental results from Haas \& Sturtevant 1987\cite{SturtevantB:87a} (symbols) with observable results (lines). \textcolor{Mpurple}{\bf \footnotesize $\pmb\times$}: incident shock, \textcolor{Mblue}{\bf \tiny $\pmb{+}$}: upstream edge of the bubble, \textcolor{Mred}{\bf \scriptsize $\pmb\Diamond$}: downstream edge of the bubble, \textcolor{Myellow}{\bf \scriptsize $\pmb\vartriangle$}: refracted shock, \textcolor{Mgreen}{\bf \scriptsize $\pmb\convolution$}: location of where the original diffracted shocks meet as shown in the schematic.}
\caption{ Cross validation estimates of the prediction error. The error bars give the $\pm$ one standard deviation of the prediction error for each joint with $R$ latent forces fitting the MLFM with only data for that joint. The horizontal `\dashed' line gives the prediction error using the combined model with $3$ forces and `\full' gives the result using $4$ forces.}
\caption{\small Survey results: The first column is the image. The second column represents the word cloud of keyword responses (responses against the reason field as shown in survey form Fig. \ref{survey}) and the third column consists of participant responses for a group's cohesion score. (Colour code for 3$^{rd}$ column: {\color{green}green}= strongly agree, {\color{blue}blue}= agree, {\color{yellow}yellow}= disagree and {\color{red}red}= strongly disagree) The fourth column shows the model prediction along with ground truth label for these images. For the 4$^{th}$ column {\color{blue}blue}= face-level prediction, {\color{red}red}= image-level prediction and {\color{orange}orange}= ground truth label). Prediction results are in the range [0 3]. In the results, the face-level network predicts the level of cohesion on the basis of emotion intensity similarity (e.g. it detects smile faces across image 2 and thus it predicts it as high cohesion). Similarly, it can not predict correctly in case of 2$^{nd}$ and 4$^{th}$ image. [Best viewed in colour]}