2015

1 — 1501.00050

\caption{Time series of the $y$-position of the leading and trailing edges of both wings for $f=23$Hz and (a) $\varphi=0$ and (b) $1.3\pi$. The leading-edge-trailing-edge phase lag $\gamma$ is shown schematically. In the $\varphi=0$ case, $\gamma_1=\gamma_2\equiv\gamma$. The $y$-axis is rendered dimensionless using $A$, the peak-to-peak amplitude swept by the leading edge at mid-chord. (c) Trailing-edge-leading-edge phase lag $\gamma$ for different forewing-hindwing phase lags. The two different markers correspond to the forewing $\gamma_1$ (\textcolor{red}{$\bigtriangleup$}) and hindwing $\gamma_2$ (\textcolor{blue}{$\bigtriangledown$}).}

2 — 1501.00909

\caption{The performance of BING- or ADOBING-enhanced trackers and their corresponding base trackers in term of AUC on the CVPR2013 benchmark. The \textcolor{red}{\bf best} and \textcolor{blue}{second best} for each tracker are indicated by \textcolor{red}{\bf red} and \textcolor{blue}{blue} respectively.}

\caption{The performance of BING- or ADOBING- enhanced trackers and their corresponding base trackers in term of \emph{success rate} on PTB. The \textcolor{red}{\bf best} and \textcolor{blue}{second best} for each tracker are indicated by \textcolor{red}{\bf red} and \textcolor{blue}{blue} respectively.}

3 — 1501.02210

\caption{Convergence of free energy as a function of the number of points in the discretization. Blue dots ({\color{blue}$\bullet$}): Legendre transform. Green triangles: ({\color{green}$\vartriangle$}) squared gradient. Black squares: ($\square$) nested loops. Red crosses ({\color{red}$\times$}): generalized scalar functionals. Simple minimizer without external derivative information. Only the Legendre method reliably converges for all $N$.}

\caption{{Derivatives of the free energy for the virus model for different discretizations and algorithms on exit from the optimization loop.} Blue dots ({\color{blue}$\bullet$}): Legendre transform, quasi-Newton. Cyan circles ({\color{cyan}$\circ$}): Legendre transform, trust-region. Red cross {\color{red}$\times$}: scalar functional, quasi-Newton. Magenta stars ({\color{magenta}$\star$}): scalar functional, trust-region.}

4 — 1501.02251

\caption{F160W surface brightness profiles of BCG+ICL as a function of radius in dlog(r[kpc])=0.15 bins, out to the point where the error in \red\surface brightness for each cluster is$<$0.2 \sbu and r$<$300 kpc. The top panel shows the observed profiles with no passband or evolution correction. The middle figure takes the observed profiles and corrects for cosmological dimming. Finally, the bottom panel accounts for evolution and passband (e+k) corrections, assuming a BC03 model, as well as cosmological dimming.}

\caption{The same as in Figure \ref{fig:f160sb}, but for \blue.}

\caption{F110W-F160W observed colour profiles, with no passband or evolution correction applied, for all four clusters in order of increasing redshift, top to bottom. We produce 2 averaged colour profiles, one for each blue filter epoch of data (solid and dashed line, respectively). All radii with $\mu_{F160W}<26$ \sbu\are plotted with error bars. ICL measurements are sensitive to background subtraction and thus our largest source of systematic uncertainty is from sky brightness variation among epochs of data. Differences between filters and observation dates drive the larger error bars and differences in colour seen in the colour profiles at large radii.}

\caption{The observed F110W-F160W averaged colour profile of each cluster, arranged by increasing slope, top to bottom. The large points are the colours of each cluster in dlog(r[kpc])=0.15 with error bars representing the standard deviation of the mean colour. Clusters are offset in colour by the amount given at the right side of the figure and the best-fit linear equations resulting from fitting the dlog(r[kpc])=0.01 profiles for each cluster are overlaid. All clusters except \MACSohthree\trend toward significantly bluer colours at higher radius with a colour gradient\colorgrad\between -0.063 and -0.101 mag log(kpc)$^{-1}$ (see Table \ref{table:fact}).}

\caption{The \red-\blue\colour profiles after passband and evolution correction with a BC03 model for a solar metallicity stellar population formed at\zform=3. Each cluster is slightly offset in radius for clarity. The large black points are the mean in each dlog(r[kpc])=0.15 bin with the standard error in the mean as the 1$\sigma$ error bars. The black line represents the linear fit found for the mean colour vs. radius relationship. The RMS scatter between the colour profiles of the four clusters is shown in the lower panel, which is low and roughly uniform from 10$<$r$<$110 kpc.}

5 — 1501.02530

\caption{Audio descriptions (DVS - descriptive video service), movie scripts (scripts) from the movies ``Harry Potter and the prisoner of azkaban'', ``This is 40'', ``Les Miserables''. Typical mistakes contained in scripts marked with \redtext{red italic}.}

6 — 1501.03139

\caption{Overview of the {\protbox} deployment architecture. The {\protbox} synchronization will take place upon {\PPairs}, which are pairs formed by one {\SharedF} and a {\ProtF}.}

7 — 1501.03379

\caption{Simulation study (Genz functions): Each panel represents one test function. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated from a shifted and scrambled Halton sequence. A Wendland regression kernel was used with $k=1$. }

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$. }

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$.}

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$.}

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$.}

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$.}

\caption{Numerical results: Each panel represents one of the 6 QMC+CF formulations. Solid lines correspond to standard QMC, dashed lines correspond to QMC+CF. $\ocircle$ represents dimension $d=1$, \textcolor{blue}{$\triangle$} represents $d=2$, \textcolor{red}{$\square$} represents $d = 3$ and \textcolor{green}{$*$} represents $d = 4$. Experiments were replicated with 10 random seeds and error bars denote standard error of the replicate mean. QMC points were generated either from a scrambled Halton sequence or a scrambled Sobol sequence (see the Main Text). The Wendland regression kernel took parameter $k$.}

8 — 1501.04158

\caption{Comparison of $F_1$-Scores between different layers and other state-of-the-art methods. $\dagger\,$We established in \cite{Neubert14} that FAB-MAP fails on the Nordland dataset. The maximum measured recall was 0.025 at 0.08 precision. $\ddagger\,$\cite{Glover08} reported that FAB-MAP is not able to address the challenges of the St. Lucia dataset, creating too many false positive matches. \textasteriskcentered For SeqSLAM on the Nordland dataset we skipped every second frame (due to performance reasons) and used a sequence length of 5, resulting in an effective sequence length of 10. All other trials with SeqSLAM used a sequence length of 10.}

9 — 1501.04378

\caption{Average Center Location Error (in pixel). Top two results are shown in \red{Red} and \blue{Blue} fonts.}

\caption{Average Overlap Rate. Top two results are shown in \red{Red} and \blue{Blue}.}

10 — 1501.04483

\caption{Normalised statistically steady-state particle concentration distributed in the wall-normal direction of the turbulent channel. \dotted $St^+=0$, \dashed $St^+=1$, \solid $St^+=10$, \dotdashed $St^+=100$; $\square$ data by \citet{sardina_schlatter_brandt_picano_casciola_2012} with the pseudo-spectral code SIMSON \citep{chevalier_schlatter_lundbladh_henningson_2007}; \textbullet~present \texttt{Nek5000} simulation.}

\caption{Mean wall-shear stress components normalised with the mean vectorial wall-shear stress; \emph{(a)} mildly curved pipe; \emph{(b)} strongly curved pipe. \dashed $\langle \tau_{w,s} \rangle$ \dotdashed $\langle \tau_{w,\theta} \rangle$ \solid $\surd{(\langle \tau_{w,s} \rangle^2+ \langle \tau_{w,\theta}\rangle^2)}$.}

\caption{Instantaneous particle concentration $C$ in the viscous sublayer ($y^+\lesssim 5$) normalised with the mean particle concentration in the domain as a function of time (convective time units): \emph{(a)} $\kappa=0.0$ \emph{(b)} $\kappa=0.01$ ,and \emph{(c)} $\kappa=0.1$. ${\blacksquare}$ \textbf{Stp0}, $\bullet$ \textbf{Stp1}, \dashed \textbf{Stp5}, \dotted \textbf{Stp10}, {\color{light-gray} \solid} \textbf{Stp25}, \solid \textbf{Stp50}, \dotdashed \textbf{Stp100}.}

\caption{\emph{(a)} Wall-normal mean particle concentration profiles for various populations (normalised by the mean particle concentration in the pipe); ${\blacksquare}$ \textbf{Stp0}, $\bullet$ \textbf{Stp1}, \dashed \textbf{Stp5}, \dotted \textbf{Stp10}, {\color{light-gray} \solid} \textbf{Stp25}, \solid \textbf{Stp50}, \dotdashed \textbf{Stp100}. \emph{(b)} The exponent of the power-law dependence $C \sim (y^+)^\alpha$ computed in the range of $1<y^+<5$ (straight pipe).}

\caption{Azimuthal distribution of mean normalised particle concentration close to the wall ($y^+ \lesssim 5$) of bent pipes with $\kappa=0.01$ \emph{(a)} and $\kappa=0.1$ \emph{(b)}. $+$ \textbf{Stp0}, \dotted \textbf{Stp5}, \dashed \textbf{Stp25}, \solid \textbf{Stp50}, and \dotdashed \textbf{Stp100}. The inner side and the outer bend of the curved pipes are located at $\theta/pi=3/2$ and $\theta/pi=1/2$, respectively.}

\caption{Distribution of the mean normalised particle concentration along the equatorial mid-plane (vertical cut) of the curved pipe with $\kappa=0.01$ \emph{(a)} and $\kappa=0.1$ \emph{(b)}. $+$ \textbf{Stp0}, \dotted \textbf{Stp5}, \dashed \textbf{Stp25}, \solid \textbf{Stp50}, and \dotdashed \textbf{Stp100}. The inner bend is located at $r/R_a=-1$ and the outer side of the curved pipe is at $r/R_a=1$}

11 — 1501.04735

\caption{(color online) (a) Typical time evolution of the plasma current with RE current. The average current decay rates $\langle dI_p/dt \rangle$ at the CQ and the RE plateau stages are determined by fitting with a linear function $I_p(t)=a+bt$. Symbol \textcolor{red}{$\odot$} corresponds to the plasma current $I_p^{(RE)}$ at the initial stage of the RE plateau. (b) The decay rates $|\langle dI_p/dt \rangle|$ versus $I_p^{(RE)}$. Symbols \textcolor{red}{$\boxdot$} correspond to the CQ rate (l.h.s. axis), and \textcolor{blue}{$\odot$} $-$ the RE plateau (r.h.s. axis). }

12 — 1501.05796

\caption{(Color online) Experimental set-up and sample a.) Schematic of the experimental setup containing a network analyzer and a waveguide. b.) Orientation of sample to electromagnetic waves in waveguide. c.) SEM\images of the JJ meta-atom: above, image of the tunnel junction which is shown in the circular indentation in the image and below, two images of a single JJ\meta-atom with 4 JJ per loop, left image, in black and white, and right image, in color. d.) SEM\image of the JJ array showing only a small fraction,\symbol{126}80 meta-atoms, of the sample.}

\caption{(Color online) Tuning the resonance with temperature and rf power. Temperature dependence when the rf power is fixed at a. )-55 dBm and b.) -49 dBm for temperatures: T = 2.0 K(black), T = 3.0 K (red), T \4.0 K(green), T = 5.0 K (blue)= 6.0 K (purple), and T = 7.0 K (brown). Aggregation of temperature-dependent data at fixed power: c.) -55 dBm and d.) - 49 dBm. Each color represents a different temperature.}

13 — 1501.06228

\caption{Quantitative measurements of defect alignment. \textbf{a,} The defect nematic order parameter, $S$, decreases as a function of the MT film's retardance. The blue shaded region represents the ``noise floor'' (see Supplementary Information). Inset: The nematic order parameter of the underlying MT filaments, $S$, also decreases with the MT film's retardance. \textbf{b,} The polar defect order parameter, $P$, showing a transition between ordered and isotropic regimes as a function of particle density. Error bars indicate a 90\% confidence interval computed by bootstrap methods (see Supplementary Information for details). \textbf{c,} The two-point nematic correlation of defect orientation $C_S(r) = \left\langle \cos 2 \left( \psi(\boldsymbol{r}) - \psi(0) \right) \right\rangle$ in MT films, which shows that orientational order persists over long distances. $l_d$ indicates the mean inter-defect spacing. The magnitude of ordering falls as retardance increases. \textbf{d,} The polar correlation of defect orientation $C_P(r) = \left\langle\cos \left( \psi(\boldsymbol{r}) - \psi(0) \right)\right\rangle$ in simulations. $\sigma$ indicates the width of a single rod (see Supplementary Information). At short ranges, defects tend to point in opposing directions, but beyond the first shell of neighbors, defects are likely to be aligned in the same direction.}

14 — 1501.06233

\caption{The distribution of objects from LAMOST DR1 in galactic coordinates. A sample of 38485 A and early--F type stars from LAMOST DR1 are represented by blue points, and 3537 Am candidates by red ones. {\color{red}{(In order to see clearly, the points in the legend box are artificially enlarged.) }}}

15 — 1501.06507

\caption{Additional \aastex\symbols}

16 — 1501.06510

\caption{(Color online) Q2D coupling parameter $\beta_{1}$ for the first $hh$ subband, taken after (\ref{eq:betaQ2D}), is shown (\textcolor{blue}{solid line}), and compared with previous analytical \cite{Habib09} (\textcolor{red}{dashed line}), experimental (symbols: \textcolor{green}{$\blacksquare$} Ref.~\cite{Winkler02}, \textcolor{magenta}{$\blacktriangle$} Refs.~\cite{Grbic04,Grbic05}, {\Large \textcolor{olive}{$\circ$}} Ref.~\cite{GrbicPhD07}, \textcolor{orange}{$\Diamond$} Ref.~\cite{Grbic08}, \textcolor{violet}{$\triangledown$} Ref.~\cite{Grbic08a}, and \textcolor{cyan}{$\blacklozenge$} Ref.~\cite{Lu98}) and numerical \cite{Winkler02,BookWinkler03} results (\textcolor{maroon}{dotted line}) based on a self-consistent method.}

\caption{(Color online) Q1D coupling parameters $\alpha_{11}$ (\textcolor{blue}{solid line}) and $\alpha'_{11}$ (\textcolor{red}{dashed line}) for GaAs, taken after (\ref{eq:alpha-1d}) and (\ref{eq:alphaP-1d}) respectively. $L_w=l_{hh}=11.8$~nm. \textit{Inset}: SOI-R strength \mbox{(\textcolor{violet}{-\textbullet-})} measured in 10~nm width holes Ge/Si core/shell nanowires from Ref.~\cite{Hao10}.}

\caption{(Color online) Q1D coupling parameters $\alpha_{11}$ (\textcolor{blue}{solid line}) and $\alpha'_{11}$ (\textcolor{red}{dashed line}) for GaAs, taken after (\ref{eq:alpha-1d}) and (\ref{eq:alphaP-1d}) respectively, for typical ($N_s>10^{11}$~cm$^{-2}$, panel (a)) and low ($N_s=10^{10}$~cm$^{-2}$, panel (b)) heavy hole densities.}

17 — 1502.00524

\caption{The system (with HN) quickly captures a simple {\it ta-tschi-bum} pattern. Time (horizontal axis) is mapped versus event labels (one line each for \emph{'ta', 'tschi',} and \emph{'bum'}). Annotated labels are indicated in black below the lines. Above the horizontal lines we find events that are correctly estimated ('{\color{green}$\bullet$}'), matched to the wrong cluster ('{\color{red}$\blacksquare$}'), and unmatched ('{\color{blue}$\blacktriangle$}') due to a wrongly estimated onset.}

\caption{Cluster merging: After 38 events, two clusters ('{\color{orange}$\bullet$}', '{\color{red}$\blacksquare$}') merge into one cluster ('{\color{red}$\blacksquare$}'). The projection of the MFCC vectors (timbre representation) onto their first two principal components (\emph{above}) and the incremental clustering tree (\emph{below}) are shown.}

\caption{Creation of new clusters: After 20 and 80 sound events, new clusters ('{\color{orange}$\bullet$}', {'\color{yellow}$\blacktriangle$}') emerge on the fly. Cf.\Fig.\ref{fig:mergeHistory}.}

18 — 1502.00926

\caption{{\small The lower bound in (\ref{noisy-system}), which is depicted by small red circles (\textcolor{red}{$\circ$}), is compared asymptotically to its approximation in (\ref{eq777}). It can be observed that (\ref{eq777}) tightly approximates the lower bound in (\ref{noisy-system}).}}

19 — 1502.02295

\caption{The average charge $\langle Q\rangle$ as a function of $\theta$ for ensembles 1 and 2 (\textcolor{red}{$\CIRCLE$}) and ensembles 3 and 4 (\textcolor{blue}{$\blacksquare$}), together with linear plus cubic fits.}

20 — 1502.02464

\caption{\label{fig:s2} (a): 3D structure spectrin network constructed by JPKPM Data processing software for AFM image obtained using JPK Nanowizard\textregistered AFM. (b): 3D structure spectrin network constructed by using NanoScope Analysis 1.5 software for AFM images obtained using Dimension FastScanTM Atomic Force Microscope.}

21 — 1502.02592

\caption{\label{fig:gaps}\coloronline (a) schematic spectrum of unitary quantum walk with band spectrum (green) and discrete spectrum (red dots) and essential gaps at $\pm1$. The arrow symbolizes the action of the symmetry operators $\ph,\ch$ on quasi-energies. We are interested in the eigenvalues on the symmetry axis indicated, around which we assume an essential gap. (b) spectrum of self-adjoint Hamiltonian marked analogously.}

\caption{\coloronline Interactive version under \explorer.\\ Left: Phase diagram showing the range of $\sixR(W)$ for translation invariant split-step walks with angle parameters $\theta_1,\theta_2$. The white lines represent gap closures, hence systems not covered in our setting. Right: Chirally off-diagonal matrix element of walk parameterized by quasi-momentum. The winding number of this curve around the origin is the topological index for translation invariant systems. This bulk index coincides with the symmetry index $\sixR(W)$, which is shown explicitly for the three points A, B, C. }

22 — 1502.02679

\caption{Monthly \gray\light curve of source~A in the energy range of 0.2$-$300 GeV.}

\caption{\fermi\\gray\spectral energy distribution of source~A fit with various models (see text). Systematic errors (see Section\ref{subsec:sa}) are indicated by black bars and the statistical errors are indicated by red bars.}

23 — 1502.02707

\caption{\label{fig:pairprop}Plot of the pairing energy (upper panel) and of the effective pairing gap $\Delta_{\rm av}+\lambda_{2,LN}$ calculated for protons (\fulldiamond)and neutrons (\opencircle) at the deformation corresponding to a minimum of the potential energy.}

\caption{\label{fig:be2-126}Selected E2 transitions in the \nuc{126}{Xe} nucleus. Theoretical results with the scaling factor ($sc=1.3$), without one ($sc=1$) and experimental data \cite{2011CO07} are shown with (\opensquare),(\opentriangle) and (\fullcircle), respectively}

\caption{\label{fig:lev21}Experimental (\fullsquare) \cite{nndc0914} and theoretical (\opensquare) energy of the $2^{+}_1$ level in the \nuc{114-144}{Xe} nuclei.}

\caption{\label{fig:e2fi}Experimental (\fullcircle) \cite{nndc0914} and theoretical (\opensquare) B(E2) probability (in W.u.) for the transition $2^{+}_1 \rightarrow 0^{+}_1$ level in the \nuc{114-144}{Xe} nuclei.}

\caption{\label{fig:e2se}Experimental (\fullcircle) \cite{nndc0914} and theoretical (\opensquare) B(E2) probability (in W.u.) for the transition $4^{+}_1 \rightarrow 2^{+}_1$ level in the \nuc{114-144}{Xe} nuclei.}

24 — 1502.03226

\caption{\label{tab:pos_FS} Comparison of the absolute values of the \emph{calculated} in-plane components of the Fermi crystal momenta in units of $\pi/a$ along the different high-symmetry lines shown in Figs.~1, 2(d), 2(e), 3, 4, and 5 of Ref.~\onlinecite{Sassa-PhysRevB-2015}. Using the plots in Fig.~3 as reference, values deviating by more than $0.02\,\pi/a$ ($0.05\,\pi/a$) are marked in \blue{blue} (\red{red}).}

25 — 1502.03719

\caption{(a) Schematic of the falling sphere experiment. The diameter of the sphere is 1 cm, and the inner diameter of the crucible is 5 cm. (b) Arrhenian plot of measured viscosities for the silica-rich phase (\textcolor{YellowOrange}{$\blacklozenge$}), $1270^\circ$-barium-rich phase (\textcolor{OliveGreen}{$\blacksquare$}) and $1070^\circ$-barium-rich phase (\textcolor{blue}{$\bullet$}). \label{fig:visco}}

26 — 1502.04323

\caption{The Gr\"{u}neisen parameter and $SDP$ for $E_{2g}^{(2)}$ obtained from various experiments and calculations of graphene and graphite are listed. The values in square brackets are the corrections by Ghandour \textit{et al.} \cite{Ghandour2013}. \textcolor{blue}{$\gamma^0$(Eq.~\ref{e4}) is the value calculated for each case from Eq.~\ref{e4}, where the values of $\gamma$ and $\gamma\prime$ are the same for all the cases and the ratio of out-of-plane strain to in-plane is obtained with the approximation that graphene and graphite have the same elastic constants.} \label{t1}}

27 — 1502.05029

\caption{Yield as a function of number of PS, $\nps$, at low {\bf (A)} and high {\bf (B)} salt concentrations. Note that for these parameters yield is zero in the absence of PSs. PSs are either all LA (\textcolor{red}{$\blacksquare$}), all high affinity ({\large\textcolor{blue}{$\bullet$}} symbols), or the Combo sequence with 1 HA and $\nps$ LA PSs(\textcolor{OliveGreen}{$\blacktriangle$} symbols). For these cases, the HA PS is placed in the center of the RNA. Results from sets of simulations with the HA PS placed in the terminal position are shown as {\bf \textcolor{green}{\pentagon}} symbols. The result from simulations with the PS binding site placed in the center of the subunits is shown as a \textcolor{gold(metallic)}{$\blacklozenge$} symbol. Note that there are 20 PS binding sites in a complete capsid, so $\nps=20$ is the stoichiometric value. Snapshots illustrate the trend in dominant outcomes with increasing PS number.}

\caption{{\bf (A-F)} Radius of hydration $\rh$ as a function of simulation time steps for assembly trajectories performed at indicated parameter values, for non-cognate RNA (\textcolor{red}{$\blacksquare$} symbols) and cognate RNA ({\large \textcolor{blue}{$\bullet$}} symbols). The $\rh$ values before subunits are introduced are shown as \textcolor{gray}{$\blacktriangle$} symbols. The subunit-subunit interaction energy $\ess$ increases from left to right. In the top row (A-C), the subunit ARM charge is (+5), and the RNA length is 575 segments; in the second row (D-F), the subunit ARM charge is (+10), and the RNA length is 910 segments. {\bf (G)} Snapshots from simulations corresponding to panel (E), with non-cognate RNA on the left and cognate RNA on the right. }

\caption{\textbf{(A)} Free energy of PS binding to a single complete, trimeric PSR binding site, as a function of PS-PRS interaction well depth $\eps$. The symbols indicate measured data points and the line shows a linear fit. \textbf{(B)} Fraction of time a PS is bound to PSRs as a function of $\eps$, measured within an assembled capsid (\textcolor{red}{$\blacksquare$} symbols) and in the presence of 12 adsorbed but unassembled subunits ({\large \textcolor{blue}{$\bullet$}} symbols). The polymer has 575 segments with one PS and $\Csalt=100$mM.}

\caption{The fraction of assembly intermediates which contain the maximum number of subunit-subunit bonds (\textbf{(A,C,E,G)} ) and the entropy of the distribution of intermediate configurations (\textbf{(B,D,F,H)} ). Results are shown as a function of $\ess$ and $\Csalt$ for a non-cognate RNA \textbf{(A,B)} and the Combo PS sequence \textbf{(C,D)} . Results are shown as a function of the number of PSs $\nps$, with $\ess=2\kt$ and $\Csalt=100$mM \textbf{(E,F)} and $\ess=6\kt$ and $\Csalt=500$mM \textbf{(G,H)} . In \textbf{(E-H)} the RNA contained either $\nps$ HA PSs ({\large\textcolor{OliveGreen}{$\bullet$}} symbols) or 1 HA PS and $\nps$ LA PSs (\textcolor{blue}{$\blacktriangle$} symbols). Results for the non-cognate RNA (no PSs) are indicated by \textcolor{red}{$\blacksquare$} symbols. }

\caption{{bf (A)} Average deviation in number of interactions $\bad$ from the ground state configuration as a function of intermediate size $n$, averaged over assembly trajectories with a uniform polyelectrolyte (\textcolor{red}{$\bullet$} symbols) or the Combo PS sequence (\textcolor{blue}{$\blacksquare$} symbols). {\bf (B)} Number of clusters as a function of intermediate size $n$ averaged over assembly trajectories. Simulation parameters were $\Csalt=50mM, \ess=7\kt$, with the cognate RNA containing the Combo PS sequence 1 HA PS + 25 LA PS.}

\caption{Yield of well-formed capsids in dynamical assembly simulations as a function of length of a polyelectrolyte with no PSs ({\large \textcolor{blue}{$\bullet$}} symbols) or 1 HA + 20 LA PSs (\textcolor{red}{$\blacksquare$} symbols). Parameters are $\ess=5\kt$ for the uniform polyelectrolyte, $\ess=2\kt$ for the PS sequence, and $\Csalt=100$mM in both cases.}

28 — 1502.05139

\caption{(Color online) $T$-$H$ phase diagram for GdCo$_{1-x}$Fe$_x$AsO, (a) $x = 0$ and (b) $x = 0.05$. The various symbols denote different types of magnetic transitions determined from the $\rho(T)$, $\chi(T)$ and $C(T)$. The \textcolor[rgb]{0.50,0.50,0.50}{$\bm{\times}$} symbols are taken from Ref. 29. The dashed and solid lines are the guides to the eyes.}

29 — 1502.07018

\caption{\textsc{Rao}s for Disk~NB, as functions of nondimensional incident wavelength. Data are grouped according to incident wave amplitude: $a=$10\,mm ({$\square$}), 20\,mm (\textcolor{red}{$\circ$}) and 40\,mm (\textcolor{blue}{$\bigtriangleup$}).}

\caption{\textsc{Rao}s for Disk~B ($\times$) and Disk~NB (\textcolor{red}{$\circ$}).}

\caption{Deviation of surge \textsc{rao} from the mean, as a function of wave steepness, for incident frequency $f = 1.25$ (left-hand panel) and 1.5 (right). Results for Disk~B ($\times$) and Disk~NB (\textcolor{red}{$\circ$}). Error bars indicate deviation of the maxima/minima.}

30 — 1502.07169

\caption{Two levels of networks in a cluster: {\color{color1} connecting CPUs in the small} and {\color{color3} servers in the large}}

\caption{Data direct I/O significantly reduces the {\color{color1} memory bus traffic} for TCP compared to classic I/O}

\caption{Query plans for TPC-H query 17: (a) for local execution, (b) for distributed execution introducing {\color{color1}exchange} operators where required, and (c) optimized with {\color{color3}pre-aggregation} and {\color{color2}broadcast} where beneficial}

31 — 1502.07362

\caption{(Color online) Electronic band structure for the spin lattice in the antiferromagnetic state (left panel) and spiral phase (right panel). The corresponding points in the phase diagram are shown by the symbols {\color{blue}{$\spadesuit$}} and {\color{red}{$\clubsuit$}} in Fig.~\ref{fig:phase-diagram}. Close to the commensurate points, $k_Fa=(2n+1)\pi/2$, the AF state fully gaps the electronic system. When the lattice spacing is increased the gaps in the AF state become smaller and at a critical value of the lattice spacing the partially gapped spiral state becomes the lowest energy state.}

\caption{(Color online) Groundstate phase diagram of the model in Eq.~(\ref{eq:H}) as a function of lattice spacing $a$ and exchange coupling $J$. Shading schematically represents the wave vector $q$ of the spiral, lying between $q=0$ in the ferromagnetic (F) phase and $q=\pi/a$ in the anti-ferromagnetic (AF) phase. Solid phase boundary lines correspond to first-order transitions, double lines to second-order transitions and dots to triple points. The symbols {\color{blue}{$\spadesuit$}} and {\color{red}{$\clubsuit$}} correspond to the parameters used for the bandstructures shown in Fig.~\ref{fig:band}.}

32 — 1502.07959

\caption{Clustering of samples from the pancreas cancer study of \citet{FiedlerLeichtle+2009} using (A) the original absence-presence data and (B) the optimized binary matrix. Filled circles indicate pancreas cancer samples, empty circles healthy controls. { \color{revision} For clustering we employed Ward's agglomerative hierarchical clustering based on a Jaccard distance matrix computed using R standard functions {\tt hclust()} with {\tt method="ward.D2"} and {\tt dist()} with {\tt method="binary"}. } \label{fig:clustering}}

33 — 1502.08018

\caption{Allowed regions in $(\alpha_3, \alpha_4)$ space for {\color{fix}the self-accelerating} branch plotted by using three conditions: $M^2_{GW} > 0$, $\rho_g > 0$ and $X_{\pm} > 0$. The left panel corresponds to $X_{+}$ solution and the right panel corresponds to $X_{-}$ solution. The shaded region with horizontal-dashed-blue line correspond to the condition $M^2_{GW} > 0$. The shaded region with vertical-dashed-black line corresponds to the condition $X_{\pm} > 0$. The grey region corresponds to the condition $\rho_g > 0$.}

34 — 1503.00509

\caption{{\it Top:} Event from the MHD simulation, at point (2.964,0.908,5.841) along a cut in the $y$-direction. Magnetic fields (left) normalized by $B_{rms} = 0.24,$ and velocity fields (right) by local upstream Alfven velocity $V_A= 0.7$. {\it Bottom}: Event from {\it Wind} spacecraft data, on January 14, 2008, 13:50 hr, normalized by $B_{rms} = 2.5$ nT, $V_A=75$ km/s. Distance $x$ is normalized by $L.$ MVF components are identified as L (\textcolor{red}{red}, solid), M( \textcolor{ForestGreen}{green}, dotted), N (\textcolor{blue}{blue}, dashed).}

\caption{B-isosurface at half-maximum value 1.11 in \textcolor{GreenYellow}{yellow}. $\bf{\bB}$-lines sampled along N-direction in \textcolor{ForestGreen}{green} and L-directions in \textcolor{BrickRed}{red}. The original 1D spatial cut is the thick {\bf black} line.}

\caption{Same as Fig.~2, except for $\bar{\bB}$ rather than $\bB.$ The isosurface is for the half-maximum value $|\bar{\bB}|=0.301$. The \textcolor{blue}{{\bf blue}} vectors on the field-lines are $\bar{\bu}$ (in the local plasma frame).}

\caption{The quasi-separatrix layer $Q_\perp=32$ in \textcolor{BlueGreen}{cyan}, and its cross-section in a plane normal to the $M$-direction in \textcolor{ElectricPink}{magenta}.}

\caption{{\it Top:} Vorticity isosurfaces at the half-maximum value $\omega=69.1$ in \textcolor{green}{green}. In \textcolor{red}{red}, origin points at time $t=0$ of magnetic field at final point $(\bx_f,t_f)=(3.31, 0.083, 6.07,2.00)$. {\it Bottom:} Backward mean-square dispersion $\langle r_\perp^2(\tau)\rangle$ orthogonal to $L$-direction as \textcolor{blue}{blue} line. Reference curve $\langle r_\perp^2\rangle=8\lambda\tau$ in \textcolor{green}{green}, and $\langle r_\perp^2\rangle= 0.02\tau^{8/3}$ in \textcolor{red}{red}.}

35 — 1503.00529

\caption{{\bf Memory of population size distribution is preserved across several diversity waves. } Time course of jumps $-\log[1-P_{collapsed}(t)]$ in the logarithm of surviving populations following a collapse of a substantial population $P_{collapsed}(t)>10^{-10}$ in A) the simplified model in which at the start of each wave all populations are set equal to each other; B) our basic model. Both were simulated at $N=1000$ and $\gamma=10^{-20}$. Note that our basic model, unlike its simplified counterpart, preserves memory of population sizes distribution across several subsequent diversity waves. This is manifested in similar fractal structure of jumps sizes in waves \#2-6 shown in panel B). Colors of symbols represent the$log10$ of the number of substantial populations during the the previous wave, when a given population originated at the small size $\gamma$. Thus red dots mark populations originated at the very end of the previous wave, while yellow dots - those originated when there were two large populations left in the previous wave. Finally, green, blue, and purple dots refer to older populations in the previous wave. }

36 — 1503.00579

\caption{Magnetic field hysteresis for the sample S1 (mass $\sim 1$~mg) measured at different temperatures. \red{Note that due to the used sample holder there is no extra magnetic background signal to be subtracted. The bottom right inset shows the same data expanding the low field region without including the virgin curves for clarity. The top left inset shows the temperature dependence of the remanent moment obtained from the field hysteresis curves.}}

37 — 1503.00638

\caption{\label{f:ehist} \blue{ Distribution of rounding errors for different levels of precision reduction. Rounding errors are scaled to the maximum granularity, $\sqrt{12fs^2}$, and the histogram is normalized to integrate to unity and thus approximates the probability density function. } }

38 — 1503.00783

\caption{\small \it Example of window proposals used in our calibration technique. $\mathcal{P}$ is the set of positive windows ({\color[rgb]{0,1,0} $\Box$}) and $\mathcal{N}$ is the set of negative windows ({\color[rgb]{1,0,0} $\Box$}) in the training set. Finally, ({\color[rgb]{1,1,0} $\Box$}) indicates E-SVMs ground-truth bounding-box. A window is positive if it has an intersection-over-union $\geq 0.5$ with a ground-truth box~\cite{everingham10ijcv}.\vspace{-2mm}}

\caption{\small \it Candidate thresholds, given the scores on positive (\textcolor{blue}{+}) and negative (\textcolor{red}{-}) windows. The only thresholds worth considering according to (\ref{eq:calib_formulation}) are the ones between a negative and positive window. Of all equivalent thresholds between two window scores, we consider only the mean of the two scores.}

39 — 1503.01109

\caption{Comparison of differential halo mass functions in GR (red circles) and F6 (blue triangles) with the \citet{Sheth1999} prediction for GR, at three redshifts -- $z=0.0$ (left panel), $0.5$ (middle panel) and $1.0$ (right panel). The relative difference between the two models is plotted in the bottom panels. Haloes are identified using a FoF algorithm with linking length 0.2. As we only have one realisation, the error bars are estimated from subsampling by dividing the simulation box into eight subboxes of equal size; the difference between F6 and GR mass functions, $\Delta n_i$, was computed for each subbox $i$, and its mean value ($\langle\Delta n\rangle$) and standard deviations ($\sigma_{\Delta n}$) were obtained using the values from the 8 subboxes; the relative difference was then calculated as $\langle\Delta n\rangle/\langle n_{\rm GR}\rangle$, with the error bars obtained in the standard way of error propagation. \tcred{The vertical dashed line indicates a cut of our FoF halo catalogue at $\sim700$ particles, or $M_{\rm FoF}\sim10^{11} h^{-1}M_\odot$, for illustrative purpose, above which there is good agreement between the GR mass functions and the Sheth-Tormen fitting formulae (better than 10\%).}}

\caption{The ratios of the differential halo mass functions between $f(R)$ gravity and GR, for the same three redshifts as in Fig.~\ref{fig:hmf_FOF}. Here the halo mass is $M_{200}$, defined as the mass within the radius at which the average density 200 times the critical density. The error bars are calculated in the same way as in Fig.~\ref{fig:hmf_FOF}. \tcred{The vertical dashed line indicates roughly the smallest halo mass ($700$ particles, or $M_{200}\sim10^{11} h^{-1}M_\odot$) we have used in the analyses of this paper.}}

40 — 1503.01187

\caption{Performance comparisons between the joint and separated decompression and decoding approaches, where the number of cells is assumed at 3, the UE transmission power is 20 dBm, the fronthaul capacity is constrained by 12 bit/c.u., and the inter-cell channel gain is assumed to be -10 dB\textcolor[rgb]{1.00,0.00,0.00}{\cite{III:SPJon}}.}

41 — 1503.01296

\caption{\label{morse}Result of a numerical determination of $\vert\psi_0(x)\vert^2$ for the Morse potential $V(x)=V_0\left(e^{-2x}-2e^{-x}\right)$ with $V_0=2$ \cite{Brndiar:2002aa}. The red solid line shows the exact result, the green points are the outcome of a Monte Carlo computation based on (\ref{psi2path}), and dotted lines indicate the $\pm 10$\% band around the exact solution.} \end{minipage} \end{figure} %\end{center} \paragraph{\textbf{The relative-weight method}} % The method of Greensite and Iwasaki \cite{Greensite:1988rr} is based on a generalization of (\ref{psi2path}) to quantum field theory. The squared VWF of the pure YM theory is given by the path integral\footnote{One could insert into the path integral a factor imposing fixing to lattice temporal gauge, which sets all timelike links to $\mathbf{1}$ except on one time slice at $t\ne 0$. However, this gauge fixing is in fact not necessary.}: % \begin{equation}\label{psi2YM} \mathbf{\Psi}^2_0[U']=\displaystyle\frac{1}{Z}\int [DU]\;\prod_{\mathbf{x},i}\delta[U_i(\mathbf{x},0)-U'(\mathbf{x})]\;\exp\left(-S[U]\right), \end{equation} % The relative-weight method \cite{Greensite:1988rr} enables one to compute ratios $\mathbf{\Psi}_0^2[U^{(n)}]/\mathbf{\Psi}_0^2[U^{(m)}]$ for configurations belonging to a finite set {${\mathcal{U}}=\left\lbrace U_i^{(j)}(\mathbf{x}),j=1,2,\dots,M\right\rbrace$} using a simple procedure: One performs Monte Carlo simulations with the usual update algorithm (\textit{e.g.\/}\ heat-bath) for all spacelike links at $t\ne0$ and for timelike links. Once in a while one updates the spacelike links at $t=0$ all at once selecting one configuration from the set $\mathcal{U}$ at random, and accepts/rejects it via the Metropolis prescription. Then, for a large number of updates $N_\mathrm{tot}$, % \begin{equation}\label{psi2ratio} \frac{\mathbf{\Psi}_0^2[U^{(n)}]}{\mathbf{\Psi}_0^2[U^{(m)}]}=\lim_{N_\mathrm{tot}\to\infty}\frac{N_n}{N_m}, \end{equation} % where $N_m$ ($N_n$) denotes the number of times the $m$-th ($n$-th) configuration is accepted. To~ensure a non-negligible acceptance rate of Metropolis updates for all configurations in the set $\mathcal{U}$, they must lie close in configuration space. This limits somewhat the applicability of the method. If the VWF is assumed to be of the form $\mathbf\Psi_0[U]={\cal{N}}\exp(-R[U])$, then the measured values of $[-\log(N_n/N_\mathrm{tot})]$ should fall on a straight line with unit slope as function of $R_n\equiv R[U^{(n)}]$. An example is shown in figure \ref{prob_k_2_2_l20_c} for a set of non-abelian constant configurations (to be specified below). %\begin{center} \begin{figure}[t!] \begin{minipage}[b]{18pc} \fbox{\includegraphics[width=17.5pc]{prob_k_2_2_l20_c.eps}} \caption{\label{prob_k_2_2_l20_c}$[-\log(N_n/N_\mathrm{tot})]$ vs.\ $R_n=\mu\kappa n$ for non-abelian constant configurations with $\kappa=0.14$; $\mu$ at $\beta=2.2$ on $20^4$ lattice was $4.06 (4)$.} \end{minipage} \hspace{1.5pc}% \begin{minipage}[b]{18pc} \fbox{\includegraphics[width=17.5pc]{prob_k_2_4_l24_n.eps}} \caption{\label{prob_k_2_4_l24_n}$[-\log(N_n/N_\mathrm{tot})]$ vs.\ $x_n=\kappa n$ for NAC configurations with $\kappa=0.14$ at $\beta=2.4$ on $24^4$ lattice; the slope $\mu$ from a linear fit comes out $6.04 (2)$.} \end{minipage} \end{figure} %\end{center} \paragraph{\textbf{Direct measurement of the vacuum wave functional}} % We used the relative-weight method to calculate the squared WVF for the following two classes of lattice gauge-field configurations: % \begin{itemize} \item[I.] \textit{Non-abelian constant (NAC) configurations\/}: % \begin{equation}\label{U_NAC} {\cal U}_\mathrm{NAC}=\left\{U_k^{(n)}(x)=\sqrt{1-\left(u^{(n)}\right)^2}\;\mathbf{1}+iu^{(n)}\bm{\sigma}_k\right\}, \end{equation} % where % \begin{equation}\label{U_NAC_2} u^{(n)}=\left(\frac{\kappa}{6L^3}n\right)^{1/4},\qquad n\in\{1,2,\dots, 10\}. \end{equation} % The constant $\kappa$, regulating amplitudes of NAC configurations, is selected so that the ratio of the smallest to the largest weight within the set is not too small, at most ${\cal{O}}\left(10^{-4}\div10^{-3}\right)$. The expected dependence of relative weights on $\kappa$ and $n$ is linear: % \begin{equation}\label{U_NAC_fit} -\log\left(\frac{N_{n}}{N_\mathrm{tot}}\right)=R_n+\mbox{const.}=\kappa n\times{\mu}+\mbox{const.} \end{equation} % The data are consistent with this expectation; the constant $\mu$ at a given coupling $\beta$ is obtained as the slope of a linear fit of $[-\log(N_{n}/N_\mathrm{tot})]$ vs.\$\kappa n$, see figure \ref{prob_k_2_4_l24_n}. This constant coincides with the parameter $\mu$ that appears in the dimensional-reduction Ansatz for the VWF (\ref{PsiDR}). Predictions for NAC configurations resulting from the DR and our proposed vacuum wave functional are identical (up to parameter-naming conventions). Moreover, the dependence of $\mu$ on $\beta$ was already computed in the pioneering work of Greensite and Iwasaki~\cite{Greensite:1988rr} more than 25 years ago, albeit at rather small-size lattices, $6^4$ and $8^4$. New results from simulations on $20^4$ and $24^4$ lattices confirm their findings, \textit{cf.\/} figure 2 of \cite{Greensite:1988rr} with our figure~\ref{mu_vs_beta_NAC_c}. The strong-coupling prediction $\mu(\beta)=\beta$ is confirmed at small $\beta$, and in the weak-coupling region $\mu(\beta)$ behaves as a physical quantity with the dimension of inverse mass: % \begin{equation}\label{mu_scaling} \mu(\beta)f(\beta)=\mu_\mathrm{phys}\approx 0.0269(3), \end{equation} % where % \begin{equation}\label{f_AF} f(\beta)=\left(\frac{6\pi^2\beta}{11}\right)^\frac{51}{121}\exp\left(-\frac{3\pi^2\beta}{11}\right). \end{equation} %\begin{center} \begin{figure}[t!] \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{mu_vs_beta_NAC_c.eps}} \caption{\label{mu_vs_beta_NAC_c}The quantity $\mu$, extracted from computed weights of NAC configurations, vs.\the coupling$\beta$.} \end{minipage} \hspace{1.5pc}% \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{prob_k_a1_2p5_k015_l30_c.eps}} \caption{\label{prob_k_a1_2p5_k015_l30_c}An example plot of $[-\log(N^{(j)}_{\footnotesize\textbf{\textit{n}}}/N_\mathrm{tot})]$ vs.\${\textstyle\frac{1}{2}}(\alpha_{\footnotesize\textbf{\textit{n}}}+\gamma_{\footnotesize\textbf{\textit{n}}}j)$ for~APW configurations.} \end{minipage} \end{figure} %\end{center} % \item[II.] \textit{Abelian plane-wave (APW) configurations\/}: % \begin{equation}\label{U_APW}{\cal U}_\mathrm{APW}=\Biggl\{U_1^{(j)}(x)=\sqrt{1-\left(w^{(j)}_{\footnotesize\textbf{\textit{n}}}(x)\right)^2}\mathbf{1}+iw^{(j)}_{\footnotesize\textbf{\textit{n}}}(x)\bm{\sigma}_3,\qquad U_2^{(j)}(x)=U_3^{(j)}(x)=\mathbf{1}\Biggr\}, \end{equation} % where % \begin{equation}\label{U_APW_2} w^{(j)}_{\footnotesize\textbf{\textit{n}}}=\sqrt{\frac{\alpha_{\footnotesize\textbf{\textit{n}}}+\gamma_{\footnotesize\textbf{\textit{n}}}j}{L^3}}\cos\left(\frac{2\pi}{L}\textbf{\textit{n}}\cdot\textbf{\textit{x}}\right),\qquad \textbf{\textit{n}}=(n_1,n_2,n_3), \qquad j\in\{1,2,\dots, 10\}. \end{equation} % Amplitudes in a particular set of plane waves with wave number $\textbf{\textit{n}}$ are parametrized by a~pair $(\alpha_{\footnotesize\textbf{\textit{n}}},\gamma_{\footnotesize\textbf{\textit{n}}})$ and depend on integer $j$. For all sets, pairs of parameters were again carefully chosen so that the actions of configurations with different $j$ in the set were not much different. For APW configurations one expects: % \begin{equation}\label{U_APW_fit} -\log\left(\frac{N^{(j)}_{\footnotesize\textbf{\textit{n}}}}{N_\mathrm{tot}}\right)=R^{(j)}_{\footnotesize\textbf{\textit{n}}}+\mbox{const.}={\textstyle\frac{1}{2}}(\alpha_{\footnotesize\textbf{\textit{n}}}+\gamma_{\footnotesize\textbf{\textit{n}}}j)\times{\omega(\textbf{\textit{n}})}+\mbox{const.} \end{equation} % For a particular wave number $\textit{\textbf{n}}$, one can plot $[-\log(N^{(j)}_{\footnotesize\textbf{\textit{n}}}/N_\mathrm{tot})]$ vs.\${\textstyle\frac{1}{2}}(\alpha_{\footnotesize\textbf{\textit{n}}}+\gamma_{\footnotesize\textbf{\textit{n}}}j)$, and determine the slope $\omega(\textbf{\textit{n}})$ from a fit of the form (\ref{U_APW_fit}). The expected linear dependence was seen in all our data sets at all couplings, wave numbers, and parameter choices; an example is displayed in figure \ref{prob_k_a1_2p5_k015_l30_c}. \end{itemize} Our main goal is to compare computed relative weights of test configurations with predictions of the DR (\ref{PsiDR}) and GO (\ref{PsiGO3D}) wave functionals. As already mentioned, NAC configurations are not suitable for that purpose, they only served us as test bed of our computer code. However, for APW the DR prediction for the dependence of the extracted function $\omega(\textbf{\textit{n}})$ on the wave number~$\textbf{\textit{n}}$: % \begin{equation}\label{DR_exp}{\omega(\textbf{\textit{n}})}\sim{\mu}\;k^2(\textit{\textbf{n}}) \qquad\dots\quad\mbox{dim.\reduction} \end{equation} % differs from our form: % \begin{equation}\label{GO_exp}{\omega(\textbf{\textit{n}})}\sim\frac{k^2(\textit{\textbf{n}})}{\sqrt{k^2(\textit{\textbf{n}})+{m}^2}} \qquad\dots\quad\mbox{GO proposal}, \end{equation} % where the wave momentum $k$ above fulfils % \begin{equation}\label{k} k^2(\textit{\textbf{n}})=2\sum_i\left(1-\cos\frac{2\pi n_i}{L}\right). \end{equation} % We therefore fitted our data for $\omega(\textbf{\textit{n}})$ at each coupling $\beta$ by the following two-parameter forms: % \begin{equation}\label{fits}{\omega(\textbf{\textit{n}})}=\left\{\begin{array}{l c l}{a}+{b}k^2(\textit{\textbf{n}}) & \dots & \mbox{dim.\reduction},\\[2mm] \displaystyle\frac{{c}k^2(\textit{\textbf{n}})}{\sqrt{k^2(\textit{\textbf{n}})+{m}^2}} & \dots & \mbox{GO proposal}. \\ \end{array}\right.\end{equation} % The result at $\beta=2.5$ ($30^4$ lattice) is displayed in figure \ref{omega_2_5_l30_c} (green dashed line for DR, blue dotted one for our proposal). Both forms fit the data quite well at low plane-wave momenta, none of them is satisfactory for larger momenta. The situation at other gauge couplings is similar. % %\begin{center} \begin{figure}[t!] \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{omega_2_5_l30_c.eps}} \caption{\label{omega_2_5_l30_c}$\omega(\textbf{\textit{n}})$ vs.\$k(\textbf{\textit{n}})$ with fits of the forms (\ref{fits}) and (\ref{best}) for APW configurations.} \end{minipage} \hspace{1.5pc}% \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{NAC_vs_APW_c.eps}} \caption{\label{NAC_vs_APW_c}The combination $(2c/m) f (\beta)$ of the best fit (\ref{best}) to APW data. Also displayed are values of $\mu f (\beta)$ extracted from NAC data.} \end{minipage} \end{figure} %\end{center} % Considerable improvement at all couplings is achieved by adding another parameter $d$ to our form: % \begin{equation}\label{best}{\omega(\textbf{\textit{n}})}= \frac{{c}k^2(\textit{\textbf{n}})}{\sqrt{k^2(\textit{\textbf{n}})+{m}^2}}\left[1+{d}k(\textit{\textbf{n}})\right], \end{equation} % see the red solid line in figure \ref{omega_2_5_l30_c}. This would correspond, in the continuum limit, to adding a~term % \begin{equation}\label{addition} d\;\left(\frac{-{\cal{D}}^2-\lambda_0}{{(-{\cal{D}}^2-\lambda_0)+m^2}}\right)^{1/2} \end{equation} % to the adjoint kernel $\mathfrak{K}$ that appears in the VWF (\ref{PsiGO3D}). For constant configurations with small amplitude the DR and GO forms coincide. For consistency between our data for NAC and APW configurations, the value of $\mu_\mathrm{NAC}$ determined from sets of NAC configurations should agree with the appropriate combination $(2c_\mathrm{APW}/m_\mathrm{APW})$ of parameters obtained for abelian plane waves. Our results clearly pass this check quite success\-fully, both values converge to each other at large$\beta$, as exemplified in figure~\ref{NAC_vs_APW_c}. Finally, the parameters of the best fit (\ref{best}), $c$, $m$ and $d$, if they correspond to some physical quantities in the continuum limit, should scale correctly when multiplied by the appropriate power of the asymptotic-freedom function $f(\beta)$ in (\ref{f_AF}). While the scaling of the ratio $(2c/m)$ multiplied by $f(\beta)$ has already been seen almost perfect in figure \ref{NAC_vs_APW_c}, it is not so convincing for individual parameters $c$ and $m/f(\beta)$, though their growth in the range of $\beta=2.2 \div 2.5$ is not large (see figure \ref{cm_vs_beta_c}) and one may optimistically hope that it will level off at still higher couplings. On the other hand, the parameter $d$ multiplied by $f(\beta)$ falls down quite rapidly in the same region (figure \ref{d_vs_beta_c}). The data thus suggest that the physical value of $d$ might vanish in the continuum limit. This signals a possibility that the form (\ref{PsiGO3D}) of the VWF might be recovered in the continuum limit. % %\begin{center} \begin{figure}[t!] \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{cm_vs_beta_c.eps}} \caption{\label{cm_vs_beta_c}Parameter $c$ (red) and rescaled mass $m/f(\beta)$ (blue) of the best fit (\ref{best}) vs.\$\beta$.} \end{minipage} \hspace{1.5pc}% \begin{minipage}[t]{18pc} \fbox{\includegraphics[width=17.5pc]{d_vs_beta_c.eps}} \caption{\label{d_vs_beta_c}Rescaled parameter $d f(\beta)$ (green points) of the best fit (\ref{best}) vs.\$\beta$.} \end{minipage} \end{figure} %\end{center} % \section{Open end of the romance; roads to take} % Let me summarize what has been achieved until now, both positive (\includegraphics[height=0.5cm]{up.png}) and negative (\includegraphics[height=0.5cm]{down.png}) points: \bigskip \noindent\begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{up.png}}] We have proposed an approximate form of the SU(2) YM vacuum wave functional that looks good in $D=2+1$, somewhat worse in $(3+1)$ dimensions. \end{itemize} \end{minipage}\hspace{1.5pc}% \begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{down.png}}] In $(3+1)$ dimensions, a method of generating configurations distributed according to the proposed vacuum wave functional is not (yet?) available. \end{itemize}\end{minipage} \bigskip \noindent\begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{up.png}}] The method of Greensite and Iwasaki allows one to compute numerically (on a lattice) relative probabilities of various gauge-field configurations in the YM vacuum. \end{itemize} \end{minipage}\hspace{1.5pc}% \begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{down.png}}] This relative-weight method is only applicable to compute weights of configurations rather close in configuration space. \end{itemize}\end{minipage} \bigskip \noindent\begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{up.png}}] For non-abelian constant configurations and for long-wavelength abelian plane waves the measured probabilities are consistent with the dimensional-reduction form, and the coefficients $\mu$ for these sets agree. \end{itemize} \end{minipage}\hspace{1.5pc}% \begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{down.png}}] Neither the dimensional-reduction form, nor our proposal for the vacuum wave functional describe data satisfactorily for larger plane-wave momenta. \end{itemize}\end{minipage} \bigskip \noindent\begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{up.png}}] Numerical data for test configurations are nicely described by a natural modification of our proposal, and the correction term seems to vanish in the continuum limit. \end{itemize} \end{minipage}\hspace{1.5pc}% \begin{minipage}[t]{18pc} \begin{itemize} \item[{\includegraphics[height=0.5cm]{down.png}}] No configurations tested so far, neither non-abelian constant nor abelian plane waves, can be considered typical. They are not in any sense true representatives of fields inhabiting the YM vacuum. \end{itemize}\end{minipage} \bigskip The investigation reported in this talk could be extended in various ways: \medskip -- One should compute, by the relative-weight method, weights of more realistic, ``typical'' configurations. One can \textit{e.g.\/} generate ensembles of configurations of the full YM theory, take link matrices from a random time slice, make their Fourier de\-com\-posit\-ion, switch on/off or modify individual momentum modes, and compare the obtained momentum dependence of relative weights of configurations with that following from our, or some other, Ansatz for the VWF.\medskip -- It is most desirable to find a way of generating field configurations distributed according to (the square of) the VWF. What was possible in $D=2+1$ dimensions, is made much more complicated by the Bianchi constraint in $(3+1)$. \medskip -- If one achieved progress in the above points and gained more evidence for the proposed form of the vacuum wave functional in $(3+1)$ dimensions, a number of other questions would call for an answer, \textit{e.g.\/}: What are the dominant field configurations in the YM vacuum? Where do colour screening and $N$-ality dependence arise from? \dots \medskip One can just hope, on the crooked road towards understanding the QCD vacuum wave functional, not to run into a ``Dead End'' sign. \ack I~am~most grateful to {Jeff Greensite} for collaboration on many topics, including those covered in the present talk, and to J\'an Pi{\v{s}}\'ut, who initiated my interest in this problem many years ago. I would also like to thank organizers of\emph{DISCRETE 2014}, in particular {Joannis Papavassiliou}, the convener of its session on strongly coupled gauge theories, for inviting me to this interesting conference in solemn edifices of King's College London. I thank Ralf Hofmann for inviting me to the \textit{4th Winter Workshop on Non-Perturbative Quantum Field Theory} in Sophia Antipolis. My research is supported by the {Slovak Research and Development Agency} under Contract No.\APVV--0050--11, and by the{Slovak Grant Agency for Science}, Project VEGA No.~2/0072/13. \appendix \section{Bird's eye view of the YM theory on the lattice\footnote{For more details see \textit{e.g.\/} the textbook \cite{Gattringer:2010zz}.}}\label{A} In the compact lattice formulation, gauge fields $\mathbf{A}_\mu(x)$ are represented by link matrices $U_\mu(x)$ (see figure \ref{lattice}): % \begin{equation}\label{U} \mathbf{A}_\mu(x)=A_\mu^a(x)\mathbf{T}^a(x)\quad\longrightarrow\quad U_\mu(x)=\exp[iga\mathbf{A}_\mu(x)]. \end{equation} % The field strengths $\mathbf{F}_{\mu\nu}(x)$ are related to products $U_{\mu\nu}(x)$ of link matrices along a plaquette: % \begin{equation}\label{Umunu} \mathbf{F}_{\mu\nu}(x)=F_{\mu\nu}^a(x)\mathbf{T}^a(x)\quad\longrightarrow\quad U_{\mu\nu}(x)= U_\mu(x) U_\nu(x+\hat{\mu})U^\dagger_\mu(x+\hat{\nu}) U^\dagger_\nu(x). \end{equation} % %\begin{center} \begin{figure}[h!] %\centerline \begin{minipage}[t]{4.5pc} ~%\caption{\label{lattice}Building blocks of pure YM theory on a lattice.} \end{minipage} \fbox{{\includegraphics[height=6cm,trim=75 0 0 0,clip,angle=270]{lattice}}} \hspace{2pc}% \begin{minipage}[t]{12pc} \caption{\label{lattice}Building blocks of pure Yang--Mills theory on a lattice.} \end{minipage} \end{figure} %\end{center} % \noindent The simplest form that replaces the euclidean SU($N$) YM action % \begin{equation}\label{S_E}{\cal S}=\frac{1}{2}\int d^4x\;\mbox{Tr}\;\left[\mbox{\textbf{F}}_{\mu\nu}(x)\mbox{\textbf{F}}_{\mu\nu}(x)\right]\end{equation} % is the Wilson action % \begin{equation}\label{S_W}{\cal S}_W= a^4\;\beta \sum_P\left[1-\frac{1}{N}\mbox{Re}\;\mbox{Tr}\;U_P\right]\quad\mbox{with}\quad \beta=2N/g^2. \end{equation} % Vacuum expectation values % \begin{equation}\label{VEV} \langle 0 \vert{\widehat{{Q}}}\vert 0\rangle = \displaystyle\int[dU]\;\mathbf{\Psi}^*_0[U]\;\widehat{{Q}}\;\mathbf{\Psi}_0[U] \end{equation} % are given by path integrals % \begin{equation}\label{VEV_PI} \langle{\widehat{{Q}}}\rangle = \displaystyle\frac{1}{Z}\int[dU_\mu(x)]\;{{Q}}[U]\;\exp(-{\cal{S}}_W[U]). \end{equation} % In the numerical Monte Carlo simulation one computes in fact: % \begin{equation}\label{VEV_MC} \langle{\widehat{{Q}}}\rangle \approx \displaystyle\frac{1}{N_\mathrm{conf}}\sum_{i=1}^{N_\mathrm{conf}}Q[\{{\cal{C}}_i\}], \end{equation} % an average over a (large) number $N_\mathrm{conf}$ of gauge-field configurations $\{{\cal{C}}_i\}$ distributed according to the probability distribution $\sim \exp(-{\cal{S}}_W[U])\sim\vert\mathbf{\Psi}_0[U]\vert^2$. \section{Recursion method for simulation of the vacuum wave functional}\label{B} % Let us define a probability distribution for gauge fields $\mathbf{A}$ in a background of a second, independent configuration $\mathbf{A}'$: % \begin{equation}\label{probability}{\cal P}[\mathbf{A};\mathfrak{K}[\mathbf{A}']]={\cal N}\left[\mathbf{A}'\right]\exp\left[-\displaystyle\int d^2x\;d^2y\; B^a(x;\mathbf{A})\;{\mathfrak{K}_{xy}^{ab}[\mathbf{A}']}\;B^b(y;\mathbf{A})\right]. \end{equation} % where $\mathbf{A}$ and $\mathbf{A}'$ are fixed to a variant of axial gauge (Appendix B of~\cite{Greensite:2007ij}). If we assume that the variance of $\mathfrak{K}$ is small among thermalized configurations, we can approximate: % \begin{equation}\label{hypothesis}{P[\mathbf{A}]}\equiv{\cal P}[\mathbf{A};\mathfrak{K}[\mathbf{A}]]\approx{\cal P}[\mathbf{A};\langle \mathfrak{K}\rangle] \approx\int d\mathbf{A}'\;{\cal P}[\mathbf{A};\mathfrak{K}[\mathbf{A}']]\;{P[\mathbf{A}']}. \end{equation} % Then the probability distribution can be generated by solving (\ref{hypothesis}) iteratively: % \begin{eqnarray} P^{(1)}[\mathbf{A}]&=&{\cal P}\left[\mathbf{A};\mathfrak{K}[0]\right],\\ P^{(k+1)}[\mathbf{A}]&=&\int d\mathbf{A}'\;{\cal P}\left[\mathbf{A};\mathfrak{K}[\mathbf{A}']\right]\; P^{(k)}[\mathbf{A}']. \end{eqnarray} % A block diagram of a practical implementation of this recursion procedure is shown below:\\ % %\centerline{{\fbox{\includegraphics[trim={-10 -10 -10 -10},width=18pc]{recursion.eps}}}} %\begin{center} \centerline{{\fbox{\includegraphics[trim={-10 -10 -10 -10}]{recursion.eps}}}} %\end{center} \bigskip\noindent In the case of the kernel appearing in the vacuum wave functional (\ref{PsiGO}), the procedure converges quite rapidly, typically in ${\cal{O}}(10)$ above cycles. The hypothesis about small changes of $\mathfrak{K}$ among equilibrated configurations is confirmed \textit{a posteriori\/} by the absence of large fluctuations of the spectrum of $\mathfrak{K}$ for individual recursion lattices. %\section*{References} \raggedright %\bibliography{DISCRETE2014} \input{vwf35.bbl} \end{document} }

42 — 1503.01347

\caption{\label{fig:kumul23}{\it Variance and skewness of the cross-spectra estimators.} The plot shows the second order cumulants, $\kappa_2$ (upper red and blue curves) and third order $\kappa_3$ (lower red and blue curves) of the PCL estimator. The blue lines correspond to results with the early reionization scenario (model 1) as fiducial model, the red when simulations has the late reionization scenario as input cosmology (model 2). The points refer to the cumulants reconstructed from the \hfic$\times$\hficq\simulations. The dashed lines correspond to the analytic full-sky derived expressions Eqs.~\ref{eq:varcross}--\ref{eq:skewcross} with a rescaled $N=(2\ell+1)\fsky$ factor. The solid lines refer to the parametrization of the cumulants based on the Edgeworth expansion defined in \refeq{cucu}.}

\caption{The offset functions $o^{A\times B}_\ell$ of the oHL likelihood for the \hfi$\times$\hficq\B-modes cross-spectra (solid) and the E-modes cross-spectra (dashed). The different colors refers to the different fiducial models used in the simulations: black is the early reionization scenario without tensor modes (model 1), red is the late reionization scenario with tensors (model 2) and blue is the\planck\2015 best fit with$r=0.1$.}

\caption{\label{fig:compvar} Ratio of the variance of the $A-B$ coss spectrum estimator to the one from the auto-spectrum of the optimally combined map. The colors indicate different noise combinations according to the following scheme: \colorcode. The dashed line recalls that most of the interesting information in EE about reinization is contained below $\ell \lesssim 15$.}

\caption{\label{fig:compskew} Ratio of the third order central moment ($\kappa_3$) for the $A\times B$ coss spectrum estimator to the one from the auto-spectrum of the optimally combined map. The colors indicate different noise combinations according to the following scheme: \colorcode. The dashed line recalls that most of the intresting information in EE about reoinization is contained below $\ell \lesssim 15$.}

43 — 1503.01375

\caption{Random orthogonal examples. We create 100 random instances and run the method 10 times per instance for a total of 1000 runs. For each metric, we report \textcolor{color1}{the total number of runs where the metric meets the desired criteria}, \textcolor{color2}{the number of instances where at least one run meets the desired criteria}, and \textcolor{color3}{the mean value of the metric}.}

\caption{Random non-orthogonal examples. We create 100 random instances and run the method 10 times per instance for a total of 1000 runs. For each metric, we report \textcolor{color1}{the total number of runs where the metric meets the desired criteria}, \textcolor{color2}{the number of instances where at least one run meets the desired criteria}, and \textcolor{color3}{the mean value of the metric for all successful attempts}. For the p.s.d.\@$\M{C}$ column, the final number is the mean number of attempts needed to find a p.s.d.\@$\M{C}$ whenever it occurs.}

44 — 1503.01407

\caption{Linear trajectory experiment: IEKF (\textcolor[rgb]{0,0,1}{\protect\rule[0.25ex]{0.5cm}{2pt}}), MEKF (\textcolor[rgb]{1,0,0}{\protect\rule[0.25ex]{0.5cm}{2pt}}), Ground truth (\textcolor[rgb]{0,0,0}{\protect\rule[0.25ex]{0.5cm}{2pt}})}

\caption{Circular trajectory experiment: IEKF (\textcolor[rgb]{0,0,1}{\protect\rule[0.25ex]{0.5cm}{2pt}}), MEKF (\textcolor[rgb]{1,0,0}{\protect\rule[0.25ex]{0.5cm}{2pt}}), Ground truth (\textcolor[rgb]{0,0,0}{\protect\rule[0.25ex]{0.5cm}{2pt}})}

45 — 1503.01459

\caption{Azimuthal displacement produced at a) L(0,3), b) F(1,2), c) L(0,4) and d) F(1,3) cutoff frequencies. Black (---), red dash (\textcolor{red}{- - -}) and blue (\textcolor{blue}{- - -}) dash lines indicate $u_r$, $u_\theta$ and $u_z$, respectively. (color online)}

46 — 1503.01678

\caption{Forecasts of SFI data set A using Lorenz's method of analogues in (a) full and (b) 2D reconstructions of the state-space dynamics. Blue {\color{blue}o}s are the true continuation $c_j$ of the time series and red {\color{red}x}s ($p_j$) are the forecasts; black error bars are provided if there is a discrepancy between the two. Embedding parameter values were estimated using standard techniques: the first minimum of the average mutual information \cite{fraser-swinney} for the delay in both images and the false near neighbor method of Kennel {\sl et al.}\cite{KBA92}, with a threshold of 20\%, for the dimension in the left-hand image. Even though the $2D$ embedding used in (b) is not faithful to the underlying topology, it enables successful forecasting of the time series.}

\caption{5,000-point forecast using the reduced-order forecast method \roLMA. Blue circles and red {\color{red}$\times$}s are the true and predicted values, respectively; vertical bars show where these values differ.}

47 — 1503.01701

\caption{(Color online) N3L, lower branch solution in a pipe. From left to right: actual solution; projection onto five response modes per Fourier mode (containing 98\% of the fluctuation energy); projection onto one response mode per Fourier mode (containing 95\% of the fluctuation energy); mean velocity profile. The red and blue shading indicates streamwise velocity fluctuation faster and slower than the mean velocity, respectively (as a fraction of the maximum amplitude streamwise velocity). The quiver arrows indicate in-plane velocity. The wall-normal region where the phase velocity is closest to the mean velocity is indicated by a dashed green line in the pipe cross-sections and a red dot indicates the phase velocity in the mean velocity profile. The lower branch solutions such as this one are close to laminar, as seen from the mean velocity profile. The mean velocity profile (---) has superposed the mean flow generated by the projections with 1, 5, 10 singular values (light to dark $-~-$) and the laminar (\textcolor{cyan}{$- \cdot$}).}

48 — 1503.03052

\caption{Methyl bisulfate molecule rotating under the action of the rotation tensor $\mathsf{R}\left(\boldsymbol{\omega}\right)$ around the orientation axis $\boldsymbol{\omega}$. The vector $\boldsymbol{R}_{\mu}$ and $\boldsymbol{R}^{\prime\prime}_{\mu}$ denote respectively the position and the ``rest'' position of the nucleus $\mu$. The vectors $\boldsymbol{\mathcal{Q}}$ and $\boldsymbol{\mathcal{P}}$ denote the position and momentum of the molecular centre of mass.} \label{frame} \end{center} \end{figure} % Applying the correspondence principle, the position, momentum and angular momentum observables associated to the center of mass are respectively given by the self-adjoint operators, % \begin{equation}\label{3.3} \begin{split} &\boldsymbol{\mathcal{Q}}=\frac{1}{\mathcal{M}}\left(\ \sum_{\mu=1}^{N} M_{\mu}\,\boldsymbol{R}_{\mu}\otimes\mathbb{1}_{e}+\sum_{\nu=1}^{n}\,\mathbb{1}_{\mathcal{N}}\otimes m\,\boldsymbol{r}_{\nu}\ \right)\\ &\boldsymbol{\mathcal{P}}=\sum_{\mu=1}^{N}\boldsymbol{P}_{\mu}\otimes\mathbb{1}_{e}+\sum_{\nu=1}^{n}\,\mathbb{1}_{\mathcal{N}}\otimes\boldsymbol{p}_{\nu}\,, \end{split} \end{equation} % where $M_{\mu}$ is the mass of the nucleus $\mu$, $m$ is the mass of the electron, $\mathcal{M}= M + nm$ is the mass of the molecule defined in terms of the mass $M$ of the nuclei. The definitions~\eqref{3.3} and the commutation relations~\eqref{3.0} imply that the operators $\boldsymbol{\mathcal{P}}$ and $\boldsymbol{\mathcal{Q}}$ satisfy the commutation relation, % \begin{equation} \label{3.6} \left[\ \boldsymbol{e}_{j}\cdot\boldsymbol{\mathcal{P}}, \boldsymbol{e}^{k}\cdot\boldsymbol{\mathcal{Q}}\ \right]= -\,i\hbar\left(\boldsymbol{e}_{j}\cdot\boldsymbol{e}^{k}\right)\mathbb{1}\,. \end{equation} % The ``relative'' position operators $\boldsymbol{R}^{\prime}_{\mu}$ and $\boldsymbol{r}^{\prime}_{\nu}$, and the ``relative'' momentum operators $\boldsymbol{P}^{\prime}_{\mu}$ and $\boldsymbol{p}^{\prime}_{\nu}$ are related respectively to the operators $\boldsymbol{R}_{\mu}$, $\boldsymbol{r}_{\nu}$, $\boldsymbol{P}_{\mu}$ and $\boldsymbol{p}_{\nu}$ by % \begin{equation}\label{3.9} \begin{split} &\boldsymbol{R}^{\prime}_{\mu}=\boldsymbol{R}_{\mu}\otimes\mathbb{1}_{e}-\boldsymbol{\mathcal{Q}}\,,\vphantom{\frac{m}{\mathcal{M}}}\\ &\boldsymbol{r}^{\prime}_{\nu}=\mathbb{1}_{\mathcal{N}}\otimes\boldsymbol{r}_{\nu}-\boldsymbol{\mathcal{Q}}\,,\vphantom{\frac{m}{\mathcal{M}}}\\ &\boldsymbol{P}^{\prime}_{\mu}=\boldsymbol{P}_{\mu}\otimes\mathbb{1}_{e}-\frac{M_{\mu}}{\mathcal{M}}\ \boldsymbol{\mathcal{P}}\,,\\ &\boldsymbol{p}^{\prime}_{\nu}=\mathbb{1}_{\mathcal{N}}\otimes\boldsymbol{p}_{\nu}-\frac{m}{\mathcal{M}}\ \boldsymbol{\mathcal{P}}\,. \end{split} \end{equation} % and satisfy the conditions, % \begin{equation}\label{3.11} \begin{split} &\sum_{\mu=1}^{N}\,M_{\mu}\,\boldsymbol{R}^{\prime}_{\mu}+\sum_{\nu=1}^{n} m\ \boldsymbol{r}^{\prime}_{\nu} = \boldsymbol{0}\,,\\ &\sum_{\mu=1}^{N}\,\boldsymbol{P}^{\prime}_{\mu}+\sum_{\nu=1}^{n}\,\boldsymbol{p}^{\prime}_{\nu} = \boldsymbol{0}\,, \end{split} \end{equation} % that are a direct consequence of the definitions~\eqref{3.3} and~\eqref{3.9}. The definitions~\eqref{3.9} and the canonical commutations relations~\eqref{3.0} and~\eqref{3.6} imply that the canonical commutations relations between the ``relative'' observables are given by, % \begin{equation}\label{3.13} \begin{split} &\left[\ \boldsymbol{e}_j\cdot\boldsymbol{P}^{\prime}_{\mu},\ \boldsymbol{e}^k\cdot\boldsymbol{R}^{\prime}_{\nu}\ \right] = -\,i\hbar\left(\boldsymbol{e}_j\cdot\boldsymbol{e}^k\right)\left( \delta_{\mu\nu}-\,\frac{M_{\mu}}{\mathcal{M}} \right)\,\mathbb{1}\,,\\ &\left[\ \boldsymbol{e}_j\cdot\boldsymbol{p}^{\prime}_{\mu},\ \boldsymbol{e}^k\cdot\boldsymbol{r}^{\prime}_{\nu}\ \right] = -\,i\hbar\left(\boldsymbol{e}_j\cdot\boldsymbol{e}^k\right)\left( \delta_{\mu\nu}-\,\frac{m}{\mathcal{M}} \right)\,\mathbb{1}\,. \end{split} \end{equation} % The ``rest'' position operators $\boldsymbol{R}^{\prime\prime}_{\mu}$ and $\boldsymbol{r}^{\prime\prime}_{\nu}$, and the ``rest'' momentum operators $\boldsymbol{P}^{\prime\prime}_{\mu}$ and $\boldsymbol{p}^{\prime\prime}_{\nu}$ are related respectively to the operators ``relative'' $\boldsymbol{R}^{\prime}_{\mu}$, $\boldsymbol{r}^{\prime}_{\nu}$, $\boldsymbol{P}^{\prime}_{\mu}$ and $\boldsymbol{p}^{\prime}_{\nu}$ by % \begin{equation}\label{3.13bis} \begin{split} &\boldsymbol{e}^{j}\cdot\boldsymbol{R}^{\prime\prime}_{\mu} = \left(\boldsymbol{e}^{j}\cdot\mathsf{R}\left(\boldsymbol{\omega}\right)^{-1}\cdot\boldsymbol{e}_{k}\right)\,\left(\boldsymbol{e}^{k}\cdot\boldsymbol{R}^{\prime}_{\mu}\right)\,,\\ &\boldsymbol{e}^{j}\cdot\boldsymbol{r}^{\prime\prime}_{\nu} = \left(\boldsymbol{e}^{j}\cdot\mathsf{R}\left(\boldsymbol{\omega}\right)^{-1}\cdot\boldsymbol{e}_{k}\right)\,\left(\boldsymbol{e}^{k}\cdot\boldsymbol{r}^{\prime}_{\nu}\right)\,,\\ &\boldsymbol{e}_{j}\cdot\boldsymbol{P}^{\prime\prime}_{\mu} = \frac{1}{2}\,\left\{\ \boldsymbol{e}^{k}\cdot\mathsf{R}\left(\boldsymbol{\omega}\right)\cdot\boldsymbol{e}_{j},\ \boldsymbol{e}_k\cdot\boldsymbol{P}^{\prime}_{\mu}\ \right\}\,,\\ &\boldsymbol{e}_{j}\cdot\boldsymbol{p}^{\prime\prime}_{\nu} = \left(\boldsymbol{e}^{k}\cdot\mathsf{R}\left(\boldsymbol{\omega}\right)\cdot\boldsymbol{e}_{j}\right)\,\left(\boldsymbol{e}_{k}\cdot\boldsymbol{p}^{\prime}_{\nu}\right)\,.\vphantom{\frac{1}{2}} \end{split} \end{equation} % where $\mathsf{R}\left(\boldsymbol{\omega}\right)$ is a rotation operator, which is a tensorial operator that is a function of the pseudo-vectorial operator $\boldsymbol{\omega}$ describing the orientation of the molecular system. The brackets $\{\ ,\ \}$ in the definitions~\eqref{3.13bis} denote an anticommutator accounting for the fact that the rotation operator $\mathsf{R}\left(\boldsymbol{\omega}\right)$ does not commute with the position operator $\boldsymbol{P}^{\prime}_{\mu}$ of the nuclei. In the definitions~\eqref{3.13bis}, we used the Einstein summation convention on alternated indices. The ``rest'' observables satisfy the conditions, % \begin{equation}\label{3.11bis} \begin{split} &\sum_{\mu=1}^{N}\,M_{\mu}\,\boldsymbol{R}^{\prime\prime}_{\mu} + \sum_{\nu=1}^{n}\,m\,\boldsymbol{r}^{\prime\prime}_{\nu} = \boldsymbol{0}\,,\\ &\sum_{\mu=1}^{N}\,\boldsymbol{P}^{\prime\prime}_{\mu}+\sum_{\nu=1}^{n}\,\boldsymbol{p}^{\prime\prime}_{\nu} = \boldsymbol{0}\,, \end{split} \end{equation} % that are a direct consequence of the conditions~\eqref{3.11} and the definitions~\eqref{3.13bis}. The definitions~\eqref{3.13bis}, the commutation relations~\eqref{3.13}, the rotational group action~\eqref{3.23bis} and the relation~\eqref{3.38} imply that the canonical commutation relations between the ``rest'' observables are given by,~\cite{Brechet:2014} % \begin{align}\label{3.17quad} &\left[\ \boldsymbol{e}_j\cdot\boldsymbol{P}^{\prime\prime}_{\mu},\ \boldsymbol{e}^k\cdot\boldsymbol{R}^{\prime\prime}_{\nu}\ \right] = -\,i\hbar\left(\boldsymbol{e}_j\cdot\boldsymbol{e}^k\right)\left(\delta_{\mu\nu}-\frac{M_{\mu}}{\mathcal{M}}\right)\,\mathbb{1}\nonumber\\ &-\left[\,\boldsymbol{e}_{j}\cdot\boldsymbol{P}^{\prime\prime}_{\mu},\,\boldsymbol{e}^{\ell}\cdot\boldsymbol{\omega}\,\right]\,\boldsymbol{e}^{k}\cdot\left(\boldsymbol{n}_{(\ell)}\left(\boldsymbol{\omega}\right)\times\boldsymbol{R}^{\prime\prime}_{\nu}\right)\,,\\ &\left[\ \boldsymbol{e}_j\cdot\boldsymbol{p}^{\prime\prime}_{\mu},\ \boldsymbol{e}^k\cdot\boldsymbol{r}^{\prime\prime}_{\nu}\ \right] = -\,i\hbar\left(\boldsymbol{e}_j\cdot\boldsymbol{e}^k\right)\left(\delta_{\mu\nu}-\frac{m}{\mathcal{M}}\right)\,\mathbb{1}\,, \end{align} % where the structural differences are due to the fact that the nuclear ``rest'' momentum $\boldsymbol{P}^{\prime\prime}_{\mu}$ does not commute with the molecular orientation operator $\boldsymbol{\omega}$ whereas the electronic ``rest'' momentum $\boldsymbol{p}^{\prime\prime}_{\nu}$ does. \section{Internal observables} The ``rest'' position and momentum observables can be recast in terms of internal observables characterising the vibrational, rotational and electronic degrees of freedom of the quantum molecular system. In order to do so, we introduce the scalar operators $Q^{\alpha}$, where $\alpha = 1,..,3N-6$, characterising the deformation amplitude of the vibrational modes of the $N$ nuclei and the vectorial operators $\boldsymbol{q}_{(\nu)}$, where $\nu = 1,..,n$, related to the position of the electrons. The ``rest'' position operators $\boldsymbol{R}^{\prime\prime}_{\mu}$ and $\boldsymbol{r}^{\prime\prime}_{\mu}$ are expressed in terms of the operators $Q^{\alpha}$, $\boldsymbol{q}_{(\nu)}$ and the equilibrium configuration $\boldsymbol{R}^{(0)}_{\mu}$ of the nucleus $\mu$, i.e.~\cite{Brechet:2014} % \begin{align}\label{3.14} &\boldsymbol{R}^{\prime\prime}_{\mu} = \boldsymbol{R}^{(0)}_{\mu}\,\mathbb{1}+\frac{1}{\sqrt{M_{\mu}}}\,Q^{\alpha}\,\boldsymbol{X}_{\mu\alpha} -\,\frac{m}{M}\!\sum_{\nu,\nu^{\prime} = 1}^{n}\!A_{\nu\nu^{\prime}}\,\boldsymbol{q}_{(\nu^{\prime})}\,,\nonumber\\ &\boldsymbol{r}^{\prime\prime}_{\nu} = \sum_{\nu^{\prime} = 1}^{n}\,A_{\nu\nu^{\prime}}\,\boldsymbol{q}_{(\nu^{\prime})}\,, \end{align} % where we used Einstein's implicit summation convention for the vibrational modes $\alpha$, and the matrix elements $A_{\nu\nu^{\prime}}$ are defined as, % \begin{equation}\label{3.15.A} A_{\nu\nu^{\prime}} \equiv \delta_{\nu\nu^{\prime}} + \frac{1}{n}\,\left(\sqrt{\frac{M}{\mathcal{M}}}-\,1\right)\,. \end{equation} % Similarly, the ``rest'' momentum operators $\boldsymbol{P}^{\prime\prime}_{\mu}$ and are expressed in terms of the operators $P_{\alpha}$, $\boldsymbol{p}_{(\nu^{\prime})}$ and the angular velocity pseudo-vectorial operator $\boldsymbol{\Omega}$, i.e.~\cite{Brechet:2014} % \begin{align}\label{3.25} &\boldsymbol{P}^{\prime\prime}_{\mu} = \boldsymbol{\Omega}\times\left(M_{\mu}\,\boldsymbol{R}^{(0)}_{\mu}\right) + \sqrt{M_{\mu}}\,P_{\alpha}\,\boldsymbol{X}^{\alpha}_{\mu}\nonumber\\ &\phantom{\boldsymbol{P}^{\prime\prime}_{\mu} =}-\,\frac{M_{\mu}}{M}\!\sum_{\nu,\nu^{\prime} = 1}^{n}\!A_{\nu\nu^{\prime}}\,\boldsymbol{p}_{(\nu^{\prime})}\,,\nonumber\\ &\boldsymbol{p}^{\prime\prime}_{\nu} = \sum_{\nu^{\prime} = 1}^{n}\,A_{\nu\nu^{\prime}}\,\boldsymbol{p}_{(\nu^{\prime})}\,. \end{align} % The definition~\eqref{3.15.A} is not unique and the particular choice made here leads to the usual structure~\eqref{3.32.1} of the canonical commutation relation between the operators $\boldsymbol{q}_{(\nu)}$ and $\boldsymbol{p}_{(\nu)}$. The vector set $\{\boldsymbol{X}_{\mu\alpha}\}$ is the basis characterising the vibrational modes that is orthonormal to the dual orthonormal basis $\{\boldsymbol{X}^{\beta}_{\mu}\}$, i.e. % \begin{equation}\label{3.18} \sum_{\mu=1}^{N}\,\boldsymbol{X}_{\mu\alpha}\cdot\boldsymbol{X}^{\beta}_{\mu} = \delta^{\beta}_{\alpha}\,. \end{equation} % In order for the identities~\eqref{3.14} and~\eqref{3.25} to satisfy the conditions~\eqref{3.11bis}, we need to impose conditions on the vectors $\boldsymbol{R}^{(0)}_{\mu}$ and $\boldsymbol{X}_{\mu\alpha}$. First, we choose the origin of the coordinate system such that it coincides with the center of mass at equilibrium, i.e. % \begin{equation}\label{3.16} \sum_{\mu=1}^{N} M_{\mu}\,\boldsymbol{R}^{(0)}_{\mu} = \boldsymbol{0}\,.\\ \end{equation} % Then, we require the deformation modes of the molecule to preserve the momentum, i.e. % \begin{equation}\label{3.17} \sum_{\mu=1}^{N}\,\sqrt{M_{\mu}}\,\boldsymbol{X}_{\mu\alpha}=\boldsymbol{0}\,. \end{equation} % We also require the deformation modes of the molecule to preserve the orbital angular momentum, i.e. % \begin{equation}\label{3.17bis} \sum_{\mu=1}^{N} \sqrt{M_{\mu}}\,\left(\boldsymbol{R}^{(0)}_{\mu}\times\boldsymbol{X}_{\mu\alpha}\right)=\boldsymbol{0}\,. \end{equation} % The constraints~\eqref{3.16}-\eqref{3.17bis} are known as the Eckart conditions~\cite{Eckart:1935}. Finally, we choose the orientation of the coordinate system such that the inertia tensor of the equilibrium position of the nuclei is diagonal, i.e. % \begin{equation}\label{3.16bis} \begin{split} &\sum_{\mu=1}^{N} M_{\mu}\left(\boldsymbol{e}_j\cdot\boldsymbol{R}^{(0)}_{\mu}\right)\left(\boldsymbol{e}_k\cdot\boldsymbol{R}^{(0)}_{\mu}\right)\\ &= \sum_{\mu=1}^{N} M_{\mu}\left(\boldsymbol{e}_j\cdot\boldsymbol{e}_k\right)\left(\boldsymbol{e}_k\cdot\boldsymbol{R}^{(0)}_{\mu}\right)^2\,. \end{split} \end{equation} % The first relation~\eqref{3.13bis} and the physical constraints~\eqref{3.16} and~\eqref{3.17bis} determine the rotation operator $\mathsf{R}\left(\boldsymbol{\omega}\right)$, i.e. % \begin{equation}\label{3.39 ter} \sum_{\mu=1}^{N}\, M_{\mu}\,\boldsymbol{R}^{(0)}_{\mu}\times\left(\mathsf{R}\left(\boldsymbol{\omega}\right)^{-1}\cdot\boldsymbol{R}^{\prime}_{\mu}\right)=\boldsymbol{0}\,.\end{equation} % % \begin{figure}[htb]\hspace{3mm} \begin{center} \includegraphics[width=0.80\columnwidth]{molecule.pdf} \caption{Molecule rotating with an angular velocity $\boldsymbol{\Omega}$ around the vertical axis coinciding with the orientation $\boldsymbol{\omega}$. The vector $\boldsymbol{R}^{(0)}_{\mu}$ and the vectorial operator $\boldsymbol{R}^{\prime\prime}_{\mu}$ are respectively the equilibrium configuration and ``rest'' position of the nucleus $\mu$. The scalar $Q_{\alpha}$ is the molecular vibration mode $\alpha$.} \label{molecule} \end{center} \end{figure} % To emphasize the physical motivation behind the previous formal development, we consider the classical counterpart of a quantum molecular system illustrated on Fig~\ref{molecule}. In a classical framework, the classical counterpart of the operatorial relation~\eqref{3.39 ter} determines the rest frame of the molecular system. Moreover, the equilibrium configuration of a molecule is given by a vector set $\{\boldsymbol{R}^{(0)}_{\mu}\}$ describing the position of the nuclei. The condition~\eqref{3.16} implies that the centre of mass of the molecule coincides with the origin of the coordinate system and the condition~\eqref{3.16bis} requires the inertial tensor of this molecule to be diagonal with respect to the coordinate system. The set of orthonormal vectors $\{\boldsymbol{X}_{\mu\alpha}\}$ characterise the $3N-6$ normal deformation modes of the molecule and thus account for the vibrations. The condition~\eqref{3.17} implies that the normal deformation modes preserve the momentum of the molecule and the condition~\eqref{3.17bis} requires that these modes also to preserve the orbital angular momentum. The internal observables are described by the scalar operators $Q^{\alpha}$, $P_{\alpha}$, the vectorial operators $\boldsymbol{q}_{(\nu)}$, $\boldsymbol{p}_{(\nu)}$ and the pseudo-vectorial operators $\boldsymbol{\Omega}$ and $\boldsymbol{\omega}$. The inversion of the definitions~\eqref{3.14} and~\eqref{3.25} yields explicit expressions for the internal observables $Q^{\alpha}$, $P_{\alpha}$, $\boldsymbol{q}_{(\nu)}$ and $\boldsymbol{p}_{(\nu)}$, i.e. % \begin{equation}\label{3.27 prime} \begin{split} &Q^{\alpha} = \sum_{\mu=1}^{N}\,\sqrt{M_{\mu}}\,\boldsymbol{X}^{\alpha}_{\mu}\cdot\Big(\boldsymbol{R}^{\prime\prime}_{\mu}-\,\boldsymbol{R}^{(0)}_{\mu}\,\mathbb{1}\Big)\,,\\ &\boldsymbol{q}_{(\nu)} = \sum_{\nu^{\prime}=1}^{n}\left(\delta_{\nu\nu^{\prime}} + \frac{1}{n}\,\left(\sqrt{\frac{\mathcal{M}}{M}}-\,1\right)\right)\,\boldsymbol{r}^{\prime\prime}_{\nu^{\prime}}\,,\\ &P_{\alpha} = \sum_{\mu=1}^{N}\frac{1}{\sqrt{M_{\mu}}}\,\left(\boldsymbol{X}_{\mu\alpha}\cdot\boldsymbol{P}^{\prime\prime}_{\mu}\right)\,,\\ &\boldsymbol{p}_{(\nu)} = \sum_{\nu^{\prime}=1}^{n}\left(\delta_{\nu\nu^{\prime}} + \frac{1}{n}\,\left(\sqrt{\frac{\mathcal{M}}{M}}-\,1\right)\right)\,\boldsymbol{p}^{\prime\prime}_{\nu^{\prime}}\,.\end{split} \end{equation} % The expressions~\eqref{3.27 prime}, the commutations relations~\eqref{3.17quad} between the ``rest'' observables and the Eckart conditions~\eqref{3.16}-\eqref{3.17bis} yield the canonical commutation relations relations between the vibrational internal observables and the electronic internal observables respectively, i.e. % \begin{equation}\label{3.32.1} \begin{split} &\left[\ P_{\alpha},\Q^{\beta}\\right]= -\,i\hbar\,\delta^{\beta}_{\alpha}\,\mathbb{1}\,,\\ &\left[\ \boldsymbol{e}_{j}\cdot\boldsymbol{p}_{(\nu)},\\boldsymbol{e}^{k}\cdot\boldsymbol{q}_{(\nu^{\prime})}\\right]= -\,i\hbar\,\left(\boldsymbol{e}_{j}\cdot\boldsymbol{e}^{k}\right)\,\delta_{\nu\nu^{\prime}}\,\mathbb{1}\,.\end{split} \end{equation} % \section{Orbital angular momentum observable} The ``relative'' orbital angular momentum operator $\boldsymbol{L}^{\prime}$ is defined as, % \begin{equation}\label{3.17pet} \boldsymbol{L}^{\prime} = \sum_{\mu=1}^{N}\,\boldsymbol{R}^{\prime}_{\mu}\times\boldsymbol{P}^{\prime}_{\mu} + \sum_{\nu=1}^{n}\,\boldsymbol{r}^{\prime}_{\nu}\times\boldsymbol{p}^{\prime}_{\nu}\,,\end{equation} % and the ``rest'' orbital angular momentum operator $\boldsymbol{L}$ is defined as, % \begin{equation}\label{3.26bis} \boldsymbol{L} = \frac{1}{2}\,\sum_{\mu=1}^{N}\,\left[\ \boldsymbol{R}^{\prime\prime}_{\mu},\\boldsymbol{P}^{\prime\prime}_{\mu}\\right]_{\boldsymbol{\times}} + \frac{1}{2}\,\sum_{\nu=1}^{n}\,\left[\ \boldsymbol{r}^{\prime\prime}_{\nu},\\boldsymbol{p}^{\prime\prime}_{\nu}\\right]_{\boldsymbol{\times}}\,,\end{equation} % where we used the notation convention $\left[\ \boldsymbol{A},\ \boldsymbol{B}\ \right]_{\boldsymbol{\times}} = \boldsymbol{A}\times\boldsymbol{B} -\,\boldsymbol{B}\times\boldsymbol{A}$. In order to express ``rest'' orbital angular momentum operator $\boldsymbol{L}$ in terms of the internal observables, we introduce the the inertia tensorial operator $\mathsf{I}\left(Q^{\,\boldsymbol{.}}\right)$. The components of the operator are defined as, % \begin{equation}\label{3.31} \boldsymbol{e}_{k}\cdot\mathsf{I}\left(Q^{\,\boldsymbol{.}}\right)\cdot\boldsymbol{e}_{\ell} = \left(\boldsymbol{e}_{k}\cdot\mathsf{I}_{0}\cdot\boldsymbol{e}_{\ell}\right)\,\mathbb{1} + Q^{\alpha}\,\left(\boldsymbol{e}_{k}\cdot\mathsf{I}_{\alpha}\cdot\boldsymbol{e}_{\ell}\right)\,, \end{equation} % where the dot in the argument of the operator $\mathsf{I}\left(Q^{\,\boldsymbol{.}}\right)$ refers to all the vibrational modes. The first term on the RHS of the definition~\eqref{3.31} is required to be diagonal with respect to the rotating molecular system according to the constraint~\eqref{3.16bis}, i.e. % \begin{align}\label{3.32bis} &\boldsymbol{e}_{k}\cdot\mathsf{I}_{0}\cdot\boldsymbol{e}_{\ell} = \left(\boldsymbol{e}_{k}\cdot\mathsf{I}_{0}\cdot\boldsymbol{e}_{k}\right)\left(\boldsymbol{e}_{k}\cdot\boldsymbol{e}_{\ell}\right)\\&\phantom{\boldsymbol{e}_{k}\cdot\mathsf{I}_{0}\cdot\boldsymbol{e}_{\ell}}= \sum_{\mu=1}^{N}\,M_{\mu}\,\left({\boldsymbol{R}^{(0)}_{\mu}}^{2}-\,\left(\boldsymbol{e}_{k}\cdot\boldsymbol{R}^{(0)}_{\mu}\right)^{2}\right)\left(\boldsymbol{e}_{k}\cdot\boldsymbol{e}_{\ell}\right)\,,\nonumber \end{align} % and the second term on the RHS, i.e. % \begin{align}\label{3.33} &\boldsymbol{e}_{k}\cdot\mathsf{I}_{\alpha}\cdot\boldsymbol{e}_{\ell} = \sum_{\mu=1}^{N}\,\sqrt{M_{\mu}}\,\left(\boldsymbol{e}_{k}\times\boldsymbol{X}_{\mu\alpha}\right)\cdot\left(\boldsymbol{e}_{\ell}\times\boldsymbol{R}^{(0)}_{\mu}\right)\nonumber\\ &\phantom{\boldsymbol{e}_{k}\cdot\mathsf{I}_{\alpha}\cdot\boldsymbol{e}_{\ell}} = \boldsymbol{e}_{\ell}\cdot\mathsf{I}_{\alpha}\cdot\boldsymbol{e}_{k}\,,\end{align} % is symmetric under the condition~\eqref{3.17bis}. Using the definitions~\eqref{3.14}-\eqref{3.25} and~\eqref{3.31}-\eqref{3.33} the ``rest'' orbital angular momentum~\eqref{3.26bis} is recast as,~\cite{Brechet:2014} % \begin{align}\label{3.27} &\boldsymbol{L} = \frac{1}{2}\,\Big\{\ \mathsf{I}\left(Q^{\,\boldsymbol{.}}\right),\\boldsymbol{\Omega}\\Big\}_{\mathsmaller{\bullet}} + \frac{1}{2}\,\sum_{\mu = 1}^{N}\,\left[\ Q^{\alpha}\,\boldsymbol{X}_{\mu\alpha},\P_{\beta}\,\boldsymbol{X}^{\beta}_{\mu}\\right]_{\times}\nonumber\\ &\phantom{\boldsymbol{L} =} + \frac{1}{2}\,\sum_{\nu = 1}^{n}\,\left[\ \boldsymbol{q}_{(\nu)},\\boldsymbol{p}_{(\nu)}\\right]_{\boldsymbol{\times}}\,,\end{align} % where we used the convention $\left\{\ \boldsymbol{A},\ \boldsymbol{B}\ \right\}_{\mathsmaller{\bullet}} = \boldsymbol{A}\cdot\boldsymbol{B} + \boldsymbol{B}\cdot\boldsymbol{A}$. In expression~\eqref{3.27}, the first term on the RHS represents the molecular orbital angular momentum, the second term corresponds to the deformation orbital angular momentum and the last term is the electronic orbital angular momentum. The orbital angular momentum operator $\boldsymbol{L}$ commutes with the operators $Q^{\alpha}$, $P_{\beta}$, $\boldsymbol{q}_{(\nu)}$ and $\boldsymbol{p}_{(\nu)}$ but it does not commute with the orientation operator $\boldsymbol{\omega}$. The commutation relation between the angular velocity operator $\boldsymbol{\Omega}$ and the molecular orientation operator $\boldsymbol{\omega}$ is given by,~\cite{Brechet:2014} % \begin{equation}\label{3.33.1} \left[\ \boldsymbol{e}^{j}\cdot\boldsymbol{\Omega},\\boldsymbol{e}^k\cdot\boldsymbol{\omega}\\right]= -\,i\hbar\,\left(\boldsymbol{e}^{j}\cdot\mathsf{I}\left(Q^{\,\boldsymbol{.}}\right)^{-1}\cdot\boldsymbol{m}^{(k)}\left(\boldsymbol{\omega}\right)\right)\,. \end{equation} % The definition~\eqref{3.27} and the commutation relation~\eqref{3.33.1} yield the commutation relation between $\boldsymbol{L}$ and $\boldsymbol{\omega}$, i.e. % \begin{equation}\label{3.35.1} \left[\ \boldsymbol{L},\\boldsymbol{e}^{j}\cdot\boldsymbol{\omega}\\right]= -\,i\hbar\\boldsymbol{m}^{(j)}\left(\boldsymbol{\omega}\right)\,. \end{equation} % The property~\eqref{3.39} and the commutation relation~\eqref{3.35.1} imply that the canonical commutation relations for the quantum description of a rotation are given by, % \begin{equation}\label{3.36.1} \left[\ \boldsymbol{n}_{(j)}\left(\boldsymbol{\omega}\right)\cdot\boldsymbol{L},\\boldsymbol{e}^{k}\cdot{\boldsymbol{\omega}}\\right]= -\,i\hbar\,\left(\boldsymbol{e}_{j}\cdot\boldsymbol{e}^{k}\right)\,. \end{equation} % In order to avoid any confusion, we would like to emphasize that the canonical commutation relations~\eqref{3.36.1} involve the orientation operator rather than the phase operator. For a one-dimensional harmonic oscillator, it is well-known that the latter is not self-adjoint since it is defined in terms of the position and momentum operators obeying canonical commutation relations~\cite{Carruthers:1968,Alimov:1979}. On the contrary, since the orientation operator $\boldsymbol{\omega}$ is a real function of the position operators, it is a self-adjoint operator. \section{Heisenberg inequalities} The vibrational canonical commutation relation~\eqref{3.32.1} implies the existence of vibrational Heisenberg inequalities, i.e. % \begin{equation}\label{3.37} \Delta\,P_{\alpha}\,\Delta\,Q^{\beta} \geq \displaystyle{\frac{\hbar}{2}}\,\delta_{\alpha}^{\beta}\,.\end{equation} % Similarly, the electronic canonical commutation relation~\eqref{3.32.1} implies the existence of electronic Heisenberg inequalities, i.e. % \begin{equation}\label{3.40} \Delta\left(\boldsymbol{e}_{j}\cdot\boldsymbol{p}_{(\nu)}\right)\,\Delta\left(\boldsymbol{e}^{k}\cdot\boldsymbol{q}_{(\nu^{\prime})}\right)\geq \displaystyle{\frac{\hbar}{2}}\,\delta_{\nu\nu^{\prime}}\,\left(\boldsymbol{e}_{j}\cdot\boldsymbol{e}^{k}\right)\,. \end{equation} % Finally, the rotational canonical commutation relation~\eqref{3.36.1} implies the existence of rotational Heisenberg inequalities, i.e. % \begin{equation}\label{3.41} \Delta\left(\boldsymbol{n}_{(j)}\left(\boldsymbol{\omega}\right)\cdot\boldsymbol{L}\right)\,\Delta\left(\boldsymbol{e}^{k}\cdot\boldsymbol{\omega}\right)\geq \displaystyle{\frac{\hbar}{2}}\,\left(\boldsymbol{e}_{j}\cdot\boldsymbol{e}^{k}\right)\,. \end{equation} % In order to compute explicitly the dispersions in the rotational Heisenberg inequalities~\eqref{3.41}, cyclic boundary conditions have to be taken carefully into account~\cite{Holevo:2011}. The orbital angular momentum $\boldsymbol{L}$ and the corresponding dispersion $\Delta\left(\boldsymbol{n}_{(j)}\left(\boldsymbol{\omega}\right)\cdot\boldsymbol{L}\right)$ can be measured using a circularly polarised light beam scattered by the spinning molecules~\cite{Allen:1992} based on a recent technique involving a rotational Doppler shift~\cite{Lavery:2013}. The shift in frequency of circularly polarized light due to the scattering is proportional the angular velocity of the molecule $\boldsymbol{\Omega}$ and to the sum of the orbital angular momentum of the molecule $\boldsymbol{L}$ and the angular momentum of the light beam. The molecular orientation axis $\boldsymbol{\omega}$ and the corresponding dispersion $\Delta\left(\boldsymbol{e}^{k}\cdot\boldsymbol{\omega}\right)$ can be measured using a strong laser pulse~\cite{Stapelfeldt:2003,Viftrup:2007}. The pulse induces an electric dipole along the direction of highest polarisability of the molecules. In order to minimise the electric dipolar energy, the dipoles align with the electric field of the laser pulse thus orienting the molecules. \section{Conclusion} In order to obtain the Rotational Heisenberg Inequalities, we establish the quantum dynamics of an isolated molecular system where all the physical degrees of freedom are described by operators, including the rotational degrees of freedom that are defined by the Lie algebra of the rotation group. Since there exists no rest frame in a quantum description of a molecular system, we used algebraic relations between the position and momentum observables associated to the nuclei and electrons in order to determine the position and momentum ``rest'' observables defined with respect to the rotating molecule. Recasting the ``rest'' observables in terms of internal observables accounting for the vibrational rotational and electronic degrees of freedom leads to one canonical commutation relation for each degree of freedom. These commutation relations yield vibrational electronic and rotational Heisenberg inequalities. The Rotational Heisenberg Inequalities~\eqref{3.41} are the product of the molecular orbital angular momentum dispersion and the molecular orientation dispersion. \bibliography{references} \bibliographystyle{eplbib} \end{document}}

49 — 1503.03070

\caption{Two-dimensional plot of the linear fit defined by the multivariate normally distributed sampling fit of Equation (\ref{propto}) depicted by the solid black line, along with the plotted points of the 24 galaxy member data set. The Milky Way is depicted distinctly in \textcolor{green}{green}. The axes $[x, y]$ depict [$\tan|P|$, $(\Sigma_{H_{I}}^{max}/(M_{\sun}pc^{-2}))/\log(M_{\star}^{bulge}/M_{\sun})$], respectively. \label{2DPlot}}

\caption{Three-dimensional plot of the plane defined by the fit of Equation (\ref{Plane}) with the multivariate normally distributed sampling depicted by a translucent \textcolor{blue}{blue} meshed surface, along with the plotted points of the 24 galaxy member included data set (depicted by \textcolor{red}{red} spheres with the Milky Way in \textcolor{green}{green}). Note that the points will appear slightly darker when they are projected behind and partially obscured by the plane. The axes $[x, y, z]$ depict [$\tan|P|$, $\log(M_{\star}^{bulge}/M_{\sun})$, $\Sigma_{H_{I}}^{max}$], respectively. Left: the view has been oriented parallel to the plane. Middle: the view has been oriented at an orientation sufficient to view the face of the plane. Right: the view has been projected along an orthogonal vector above the plane.\label{3DPlot}}

50 — 1503.03245

\caption{(Color online) $\gamma$-ray de-excitation spectra associated with the Coulomb excitation of $^{202}$Rn on $^{109}$Ag at 2.90~MeV/u, Doppler-corrected for projectiles (black) and target recoils ({\color{red}red}). Only events identified in prompt coincidence with a recoiling target nucleus are shown; random events, with respect to the particle-$\gamma$ coincidence time, have been subtracted. Peaks are marked with their energy in keV.}

\caption{(Color online) $\gamma$-ray de-excitation spectra associated with the Coulomb excitation of $^{204}$Rn on $^{109}$Ag at 2.90~MeV/u, Doppler-corrected for projectiles (black) and target recoils ({\color{red}red}). Only events identified in prompt coincidence with a recoiling target nucleus are shown.; random events, with respect to the particle-$\gamma$ coincidence time, have been subtracted. Peaks are marked with their energy in keV.}

\caption{(Color online) Experimental $\langle\beta_{2}^{2}\rangle^{1/2}$ values deduced from the $B(E2; 0^{+}_{1} \to 2^{+}_{1})$ values measured in this work (black circles, ``CLX'') and those from isotope-shift measurements and liquid-drop model for both the ground ({\color{red}red} down triangles, ``IS(g)'') and isomeric ({\color{red}red} up triangles, ``IS(m)'') states. The uncertainties on the latter are dominated by the propagation of the uncertainty in the Grodzins-Raman rule~\cite{Raman2001}, which is a systematic contribution. The isotope shift values are slightly offset from integer $A$ values for clarity of presentation.}

51 — 1503.03355

\caption{Datasets analyzed \label{tab:datasets}}{ \rowcolors{1}{lightgray}{white} \begin{tabular}{llllll} & \textbf{Name} & \textbf{Description} & \textbf{Dimensions} & {\bf Number of nonzeros} \\ & \enron & (sender, recipient, month) & $186\times186\times44$ & 9838 \\ \hiderowcolors & \reality\cite{eagle2009inferring} & (person, person, means of communication) &$88\times88 \times 4$ & 5022 \\ %& LBNL \cite{pang2005first} & (src, dst, port \#) & $65170 \times 65170 \times 65327$ & 27269 \\ & \facebook \cite{viswanath-2009-activity} & (wall owner, poster, day) & $63891\times 63890 \times 1847$ & 737778 \\ %& Taxi-1000 & (latitude, longitude,minute) & $1000\times1000\times$ & \\ & \taxi\cite{yuan2011driving,Wang2014travel} & (latitude, longitude,minute) & $100\times100\times 9617$ & 17762489 \\ & \dblp \cite{papalexakis2013more} & (paper, paper, view)&$7317\times7317\times3$ & 274106\\ & \netflix & (movie, user, date) & $17770\times252474\times88$ & 50244707 \\ & \amazonco\cite{snapnets} & (product, product, product group)& $256\times256\times5$ & 5726 \\ & \amazonmeta\cite{snapnets} & (product, customer, product group) & $10000\times263011\times5$ & 441301\\ & \yelp & (user, business, term) &$43872\times11536\times10000$ & 10009860\\ & \airport & (airport, airport, airline) & $9135\times9135\times19305$ & 58443 \\ \end{tabular} }

52 — 1503.03874

\caption{Radial velocity measurements using A- and B-type stellar spectra are hindered by rotational broadening of their observed spectral features. \textcolor{black}{Here, we see this effect as illustrated by the $H\delta$ 410.1 nm Balmer line.}{\em Upper panel}: One HIRES echelle order of the Solar spectrum measured by observing reflected Sunlight from the asteroid Vesta. This spectrum is representative of those of low-mass stars observed by the CPS program. {\em Lower panel}: the same HIRES order, this time showing an observation of an A--type star, HR\,6827. This rapidly-rotating star has hundreds of times fewer spectral features than are seen in the Solar spectrum. The high-mass star cannot be analyzed in the same way as a Sun-like due to its broad spectral features, which are significant fraction of their echelle orders.}

\caption{ Three representative orders from a spectrum of HR\,8634, spanning roughly 80 angstroms each. Data points and residuals in black with the best fit transformed PHOENIX model in red. The bottom panel shows the residuals in fractional values of the total flux at each pixel. The trend in residuals around the lines are due to the imperfect fit of the theoretical line profile model to the data. The fit precision is worse for theoretical models than for relative radial velocities, which is part of why the absolute radial velocity precision is worse than the relative radial velocity precision.}

\caption{The phase-folded radial velocity time series of HR\,3067, an A-type star with a predicted companion at an orbital period of 1.5 years\citep{2012A&A...546A..69M}. The RVs show an orbit with a longer period than astrometry predicted and significant eccentricity. }

53 — 1503.03885

\caption{\sm Absolute brightness distribution of the current Mars crossing objects. Bin size is 0.5 magnitude; note that the ordinate is in logarithmic scale. The dashed line is the best-fit straight line for $14<H<15$. {\color{red} [This needs to be revised.]}}

\caption{\sm The impact flux of bright MCOs ($H<16$) on Mars and its seasonal variation. We report the average of three simulations; the standard deviations of the fluxes obtained in these simulations are all below 3.5\% of the reported flux. {\color{red}\bf[Should drop the $\pm$ uncertainties.]}}

54 — 1503.03948

\caption{ Video validation. (A) We calculated errors from video tracking using a 20cm x 20cm calibration grid moving through {\color{revision}the head motion region of space.} When the calibration grid displaced or rotated with a sagittal orientation (i.e. planar measurements), errors were always sub-millimeter. When depth measurement was involved, with the grid rotating in non-sagittal directions, errors were larger but still within 2mm. (B) We fit a t location-scale distribution to the error, and there is less than 2.5\% total probability of errors greater than 1mm. (C) We also verified that video-derived sagittal kinematics agree well with those measured by the sensors, which further confirms our video measurements. }

55 — 1503.05133

\caption{Matching a data block $\rv{B}^m = \rv{B}_1\!\ldots\!\rv{B}_m$ to output symbols $\tilde{\rv{A}}^n = \tilde{\rv{A}}_1\!\ldots\!\tilde{\rv{A}}_n$ and reconstructing the original sequence at the dematcher. The rate is $\frac{m}{n}\left[ \frac{\text{bits}}{\text{output symbol}} \right]$. The matcher can be interpreted as emulating a discrete memoryless source $\pmf{A}$.} \label{fig:memoryless_Source} \end{figure} \section{Problem statement} \label{sec:problem} The entropy of a discrete random variable $\rv{A}$ with alphabet $\mset{A}$ and distribution $\pmf{A}$ is \begin{equation} \entrp{\rv{A}} = \sum_{a \in \supp(\pmf{A})} -\pmf{A}(a) \log_2 {\pmf{A}(a)} \end{equation} where $\supp(\pmf{A}) \subseteq \mset{A}$ is the support of $\pmf{A}$. The informational divergence of two distributions on $\mset{A}$ is \begin{equation} \diverg{\pmf{\hat{A}}}{\pmf{A}} = \sum_{a \in \supp(\pmf{\hat{A}})} \pmf{\hat{A}}(a) \log_2 \frac{\pmf{\hat{A}}(a)}{\pmf{A}(a)}. \end{equation} The normalized informational divergence for length $n$ random vectors $\rv{\hat{A}}^n = \rv{\hat{A}}_1\!\ldots\!\rv{\hat{A}}_n$ and $\rv{A}^n$ is defined as \begin{equation} \frac{\diverg{\pmfn{\hat{A}}{n}}{\pmfn{A}{n}}}{n}. \end{equation} For random vectors with independent and identically distributed (iid) entries, we write \begin{equation} \pmf{A}^n(a^n) = \prod_{i=1}^n \pmf{A}(a_i). \end{equation} A one-to-one f2f distribution matcher is an invertible function $f$. We denote the inverse function by $f^{-1}$. The mapping imitates a desired distribution $\pmf{A}$ by mapping $m$ \beronehalf distributed bits $\rv{B}^m$ to length $n$ strings $\tilde{\rv{A}}^n = f(B^m) \in \mset{A}^n$. The output distribution is $\pmfn{\tilde{A}}{n}$. The concept of one-to-one f2f distribution matching is illustrated in Fig. \ref{fig:memoryless_Source}. \begin{definition} \label{def:achievableRate} A matching rate $R = m/n$ is achievable for a distribution $\pmf{A}$ if for any $\alpha > 0$ and sufficiently large $n$ there is an invertible mapping $f\colon\lbrace0,1\rbrace^m \to \mset{A}^n$ for which \begin{equation} \frac{\diverg{P_{f(\rv{B}^m)}}{\pmf{A}^n}}{n}\leq \alpha. \label{eq:ratecond:diverg} \end{equation} \end{definition} The following proposition in \cite{bocherer2014informational} relates the rate $R$ and \eqref{eq:ratecond:diverg}. \begin{proposition}[Converse,{\cite[Proposition~8]{bocherer2014informational}}] \label{prop:converse} There exists a positive-valued function $\beta$ with \begin{equation} \beta(\alpha) \overset{\alpha \rightarrow 0}{\longrightarrow} 0 \end{equation} such that \eqref{eq:ratecond:diverg} implies \begin{equation} \frac{m}{n} \leq \frac{\entrp{\rv{A}}}{\entrp{\rv{B}}} + \beta(\alpha). \end{equation} \end{proposition} Proposition \ref{prop:converse} bounds the maximum rate that can be achieved under condition \eqref{eq:ratecond:diverg}. Since $\entrp{\rv{B}} = 1$ we have \begin{equation} R \leq \entrp{\rv{A}} \label{eq:conv:result} \end{equation} for any achievable rate $R$. % The image of $f$ is $\supp(\pmfn{\tilde{A}}{n})$ and has $2^m$ elements. % We only address $2^{m}$ codewords which does not match with the request to create an independent and identical distributed (iid) process. % There are sequences that are not in the support i.e. they can never occur although an iid source would possibly create them. However there exist matcher $f$ that achieve maximum rate according to definition \ref{def:achievableRate}. This contradiction is resolved in the following remark. % \begin{remark} % \cite[Theorem 1.2]{wyner1975common} states that % an arbitrary distributed source outputs entropy typical sequences of an iid process for a large blocksize if distance (in terms of normalized divergence) between former and latter distribution is small. Thus an arbitrary distributed source imitates an iid source if normalized divergence is sufficient small. % \end{remark} \section{Constant Composition Distribution Matching} \label{sec:ConstanCompositionDistributionMatching} The empirical distribution of a vector $\boldsymbol{c}$ of length $n$ is defined as \begin{equation} P_{\mathsf{\bar{A}},\boldsymbol{c}}(a) := \frac{n_a(\boldsymbol{c})}{n} \end{equation} where $n_a(\boldsymbol{c}) = \left|\left\lbrace i: c_i = a \right\rbrace \right|$ is the number of times symbol $a$ appears in $\boldsymbol{c}$. The authors of \cite[Sec.~2.1]{csiszar2004information} call $P_{\mathsf{\bar{A}},\boldsymbol{c}}$ the \emph{type} of $\boldsymbol{c}$. An $n$-type is a type based on a length $n$ sequence. A codebook $\mset{C}_{\!\text{ccdm}} \subseteq{\mset{A}}^{n}$ is called a \emph{constant composition code} if all codewords are of the same type, i.e., $n_a(\boldsymbol{c})$ does not depend on the codeword $\boldsymbol{c}$. We will write $n_a$ in place of $n_a(\boldsymbol{c})$ for a constant composition code. %\begin{definition} %Let $n_a(c)$ be the number of symbols $a \in \mathcal{A}$ occurring in the codeword $c$. %A codebook $\mset{C}_{\!\text{ccdm}} \subseteq{\mset{A}}^{n}$ is called a constant composition codebook if and only if $n_a(\boldsymbol{c}_j) = n_a(\boldsymbol{c}_k)\quad \forall{a \in \mset{A};\boldsymbol{c}_j, \boldsymbol{c}_k \in \mset{C}}$ %\end{definition} %Since by definition $n_a(c) = n_a, \;\forall c \in \mset{C}_{\!\text{ccdm}}$ \subsection{Approach} We use a constant composition code with $n_a \approx \pmf{A} n$. As all $n_a$ need to be integers and add up to $n$, there are multiple possibilities to choose the $n_a$. We use the allocation that solves \begin{equation} \begin{split} \pmf{\bar{A}} = &\argmin_{\pmf{\bar{A}'}} \:\diverg{\pmf{\bar{A}'}}{\pmf{A}}\\ &\st\pmf{\bar{A}'} \text{ is }n\text{-type}. \end{split}\label{eq:quantization} \end{equation} The solution of \eqref{eq:quantization} can be found efficiently by \cite[Algorithm~2]{bocherer2014optimal}. Suppose the output length $n$ is fixed and that we can choose the input length $m$. Let $\nTypePA{n}{\pmf{\bar{A}}}$ be the set of vectors of type $\pmf{\bar{A}}$, i.e., we have \begin{equation} \nTypePA{n}{\pmf{\bar{A}}} =\lbrace\boldsymbol{v} \;\vert \;\boldsymbol{v} \in \mset{A}^n, \frac{n_a(\boldsymbol{v})}{n} = \pmf{\bar{A}}(a)\quad \forall a \in \mset{A}\rbrace. \end{equation} The matcher is invertible, so we need at least as many codewords as input blocks. The input blocklength must thus not exceed $ \log_2 \nTypePAcard $. We set the input length to $m = \lfloor\log_2 \nTypePAcard \rfloor$ and we define the encoding function \begin{equation} f_{\!\text{ccdm}}: \lbrace 0,1 \rbrace^m \to \nTypePA{n}{\pmf{\bar{A}}}. \end{equation} The actual mapping $f_{\!\text{ccdm}}$ can be implemented efficiently by arithmetic coding, as we will show in Section~\ref{sec:ArithmeticCoding}. The constant composition codebook is now given by the image of $f_{\!\text{ccdm}}$, i.e., \begin{equation} \mset{C}_{\!\text{ccdm}} = f_{\!\text{ccdm}}(\lbrace 0,1 \rbrace^m). \end{equation} Since $f_{\!\text{ccdm}}$ is invertible, the codebook size is $|\mset{C}_{\!\text{ccdm}}|=2^m$. \subsection{Analysis} \label{sec:analysis} We show that $f_{\!\text{ccdm}}$ asymptotically achieves all rates satisfying \eqref{eq:conv:result}. We can bound $m$ by \begin{equation} m = \left\lfloor\log_2 \nTypePAcard \right\rfloor \geq \log_2 \nTypePAcard -1. \label{eq:bound_kc} \end{equation} Recall that the matcher output distribution is $\pmfn{\tilde{A}}{n}$. We have \begin{align} &\diverg{\pmfn{\tilde{A}}{n}}{\pmf{A}^{n}} = \sum_{a^{n} \in \mathcal{C}_{\!\text{ccdm}} \subseteq \nTypePA{n}{\bar{P}_{\rv{A}}}} 2^{-m} \log_2\frac{2^{-m}}{\pmf{A}^n(a^n)}\frac{\pmf{\bar{A}}(a^n)}{\pmf{\bar{A}}(a^n)}\nonumber \displaybreak[3]\\ &\quad= \diverg{\pmfn{\tilde{A}}{n}}{\pmf{\bar{A}}^{n}} + \sum_{a^{n} \in \mathcal{C}_{\!\text{ccdm}} \subseteq \nTypePA{n}{\bar{P}_{\rv{A}}}} 2^{-m} \log_2\frac{\pmf{\bar{A}}^n(a^n)}{\pmf{A}^n(a^n)} \nonumber \displaybreak[3]\\ &\quad= \diverg{\pmfn{\tilde{A}}{n}}{\pmf{\bar{A}}^{n}} + \vert \mathcal{C}_{\!\text{ccdm}} \vert 2^{-m} \sum_{a \in \mset{A}} n_a \log_2 \frac{\pmf{\bar{A}}(a)}{\pmf{A}(a)}\nonumber \\ &\quad= \underbrace{\diverg{\pmfn{\tilde{A}}{n}}{\pmf{\bar{A}}^{n}}}_{\text{Term 1}}+ n \underbrace{\diverg{\pmf{\bar{A}}}{\pmf{A}}}_{\text{Term 2}}.\label{eq:divergenceCCC} \end{align} For Term 1 we obtain \begin{align} \diverg{\pmfn{\tilde{A}}{n}}{\pmf{\bar{A}}^{n}} &= \sum_{a^{n} \in \mathcal{C}_{\!\text{ccdm}} \subseteq \nTypePA{n}{\bar{P}_{\rv{A}}}} 2^{-m}\log_2 \frac{2^{-m}}{\prod\limits_{i \in \mset{A}}\pmf{\bar{A}}(i)^{n_i}}\notag\\ &= \sum\limits_{\mathcal{C}_{\!\text{ccdm}}} 2^{-m} \log_2 \frac{2^{-m}}{2^{-n\mathbb{H}(\rv{\bar{A}})}}\notag\\ &= n\mathbb{H}(\rv{\bar{A}})-m.\label{eq:asymptoticDivergence} \end{align} %--------------- Mark1 ------------- Using \eqref{eq:asymptoticDivergence} in \eqref{eq:divergenceCCC} and dividing by $n$ we have \begin{equation} \frac{\diverg{\pmfn{\tilde{A}}{n}}{\pmf{A}^{n}}}{n} = \mathbb{H}(\rv{\bar{A}}) - R + \diverg{\pmf{\bar{A}}}{\pmf{A}}. \label{eq:normdivergExplicit} \end{equation} The choice \eqref{eq:quantization} of $\pmf{\bar{A}}$ guarantees (see \cite[~Proposition 4]{bocherer2014optimal}) that for the third term in \eqref{eq:normdivergExplicit} we have \begin{equation} \diverg{\pmf{\bar{A}}}{\pmf{A}} < \frac{k}{\displaystyle \min_{a \in \supp\pmf{A}}\pmf{A}(a) n^2} \label{eq:upperboundOnChoiceOfPABar} \end{equation} where $k = |\mset{A}|$ is the alphabet size. Consequently, we know that this term vanishes as the blocklength approaches infinity, i.e., we have \begin{equation} \lim_{n\rightarrow \infty} \diverg{\pmf{\bar{A}}}{\pmf{A}} = 0. \label{eq:probbilityidentity} \end{equation} We now relate the input and output lengths to understand the asymptotic behavior of the rate. By \cite[~Lemma 2.2]{csiszar2004information}, we have \begin{equation} \nTypePAcard \geq \binom{n + k - 1}{k - 1}^{-1} 2^{n\mathbb{H}(\rv{\bar{A}})} \geq (n+k)^{-k} 2^{n\mathbb{H}(\rv{\bar{A}})}. \end{equation} Taking the logarithm to the base $2$ and dividing by $n$ we have \begin{equation} \frac{\log_2{\nTypePAcard}}{n} \geq \frac{-k \log_2(n+k)}{n}+ \entrp{\rv{\bar{A}}}. \label{eq:bound_log} \end{equation} For the rate, we obtain \begin{align} R = \frac{m}{n} &\overset{\eqref{eq:bound_kc}}{\geq} \frac{\log_2 \nTypePAcard}{n} - \frac{1}{n} \notag\\ &\overset{\eqref{eq:bound_log}}{\geq} \frac{-k \log_2(n+k)}{n}+ \entrp{\rv{\bar{A}}} - \frac{1}{n} \label{eq:lowerBoundOnRate} \end{align} and in the asymptotic case \begin{equation} \lim_{n \rightarrow \infty} R = \entrp{\rv{\bar{A}}}. \label{eq:asymptoticRate} \end{equation} From \eqref{eq:probbilityidentity} and \cite[Proposition~6]{bocherer2014informational} we know that $\entrp{\rv{\bar{A}}} \rightarrow \entrp{\rv{A}}$ and by \eqref{eq:probbilityidentity} and \eqref{eq:asymptoticRate} in \eqref{eq:normdivergExplicit}, normalized divergence approaches zero for $n \rightarrow \infty$. \begin{figure} \centering \input{figures/rate_diverg_blocksize_ccadm} \caption{Normalized divergence and rate of ccdm over output blocklength for $\pmf{A}=(0.0722,0.1654,0.3209,0.4415)$. For comparison, the performance of optimal f2f \cite[Sec.~4.4]{amjad2013algorithms} and aadm \cite{schulte2014Zero} is displayed. Because of limited computational resources, we could calculate the performance of optimal f2f only up to a blocklength of $n = 90$.} \label{fig:normalizedDivergenceAndRateVsBlocklength} \end{figure} \begin{example} The desired distribution is \begin{equation*} \pmf{A}=(0.0722,0.1654,0.3209,0.4415). \end{equation*} Fig.~\ref{fig:normalizedDivergenceAndRateVsBlocklength} shows the normalized divergences and rates of ccdm and the optimal f2f length matcher \cite[Sec.~4.4]{amjad2013algorithms}. The empirical performance of aadm \cite{schulte2014Zero} is also displayed. For optimal f2f and aadm, the rate is fixed to $\mathbb{H}(\rv{A})$ bits per symbol. Observe that the ccdm needs about 4 times the blocklength of the optimal scheme to reach an informational divergence of $0.06$ bits per symbol. However, the memory for storing the optimal codebook grows exponentially in $m$. For $n = 10$, we already need about $10240$ bits = $1.25$ kB; for $n = 100$ we would need $1.441\times10^{19}$ TB of memory. In this example, ccdm performs better than aadm for short blocklength up to $100$ symbols. Fig.~\ref{fig:normalizedDivergenceAndRateVsBlocklength} also shows the lower and upper bounds \eqref{eq:conv:result} and \eqref{eq:lowerBoundOnRate}, respectively. \end{example} \section{Arithmetic Coding} \label{sec:ArithmeticCoding} \begin{figure} \centering \input{diagrams/ccadm_overview} \caption{Diagram of a constant composition arithmetic encoder with $\pmf{\bar{A}}(0) = \pmf{\bar{A}}(1) = 0.5$, $m=2$ and $n=4$.} \label{fig:ccadm_overview} \end{figure} We use arithmetic coding for indexing sequences efficiently. % In data compression arithmetic coding is widely used because the data model for the input stream can be easily adapted \cite[Sec.~3.1]{Sayir1999coding}. % This way it is possible to react on changes of the data distribution. % In case of distribution matching the statistics of the input are assumed to be \beronehalf. % We are interested in controlling the output sequence. % For this reason we connect a statistical model to the output that is fed by the output symbols as shown in Figure \ref{fig:block_enc_dec>parameter_estimator}. % \begin{figure}[ht] % \centering % \input{diagrams/block_enc_dec_parameter_estimator} % \caption{Block diagram on arithmetic encoding with parameter estimation on the output symbols} % \label{fig:block_enc_dec>parameter_estimator} % \end{figure} Our arithmetic encoder associates an interval to each input sequence in $\{0,1\}^m$ and it associates an interval to each output sequence in $\nTypePA{n}{\pmf{\bar{A}}}$, see Fig.~\ref{fig:ccadm_overview} for an example. The size of an interval is equal to the probability of the corresponding sequence according to the input and output model, respectively. For the input model we choose an iid \beronehalf process. We describe the output model by a random vector \begin{equation} \rv{\bar{A}}^n = \rv{\bar{A}}_1\rv{\bar{A}}_2\ldots\rv{\bar{A}}_n \end{equation} with marginals $\pmfd{\bar{A}}{i}=\pmf{\bar{A}}$ and the uniform distribution \begin{equation*} \pmfn{\bar{A}}{n} (a^n)= \frac{1}{\nTypePAcard}\quad\forall a^n\in \nTypePA{n}{\pmf{\bar{A}}}. \end{equation*} The intervals are ordered lexicographically. All input and output intervals range from 0 to 1 because all probabilities add up to 1. \begin{example}\label{ex:arithmBasic} Fig.~\ref{fig:ccadm_overview} shows input and output intervals with output length $n=4$ and $\pmf{\bar{A}}(0) = \pmf{\bar{A}}(1) = 0.5$. There are 4 equally probable input sequences and 6 equally probable output sequences. The intervals on the input side are $[0,0.25)$, $[0.25, 0.5)$, $[0.5,0.75)$ and $[0.75,1)$. The intervals on the output side are $[0,\frac{1}{6})$, $[\frac{1}{6},\frac{2}{6})$, $[\frac{2}{6},\frac{3}{6})$, $[\frac{3}{6},\frac{4}{6})$, $[\frac{4}{6},\frac{5}{6})$ and $[\frac{5}{6},1)$.\footnote{Please note that in this case no distribution matcher is needed. However, the invertible mapping is of interest in its own right.} \end{example} The arithmetic encoder can link an output sequence to an input sequence if the lower border of the output interval is inside the input interval. In the example (Fig.~\ref{fig:ccadm_overview}) '00' may link to both '0101' and '0011', while for '01' only a link to '0110' is possible. There are at most two possible choices because by \eqref{eq:bound_kc} the input interval size is less than twice the output interval size. Both choices are valid and we can perform an inverse operation. In our implementation, the encoder decides for the output sequence with the lowest interval border. As a result, the codebook $\mset{C}_{\!\text{ccdm}}$ of Example \ref{ex:arithmBasic} is $\lbrace$'$0011$', '$0110$, '$1001$', '$1100$'$\rbrace$. In general $\mset{C}_{\!\text{ccdm}}$ has cardinality $2^m$ with $2^m \leq \nTypePAcard < 2^{m+1}$ according to \eqref{eq:bound_kc}. It is not possible to index the whole set $\nTypePA{n}{\pmf{\bar{A}}}$ unless $2^m = \nTypePAcard $. The analysis of the code (Section \ref{sec:analysis}) is valid for all codebooks $\mset{C}_{\text{\!ccdm}} \subseteq \nTypePA{n}{\pmf{\bar{A}}}$. The actual subset is implicitly defined by the arithmetic encoder. We now discuss the online algorithm that processes the input sequentially. Initially, the input interval spans from 0 to 1. As the input model is \beronehalf we split the interval into two equally sized intervals and continue with the upper interval in case the first input bit is '1'; otherwise we continue with the lower interval. After the next input bit arrives we repeat the last step. After $m$ input bits we reach a size $2^{-m}$ interval. After every refinement of the input interval the algorithm checks for a sure prefix of the output sequence, e.g., in Fig.~\ref{fig:ccadm_overview} we see that if the input starts with 1 the output must start with 1. Every time we extend the sure prefix by a new symbol, we must calculate the probability of the next symbol given the sure prefix. That means we determine the output intervals within the sure interval of the prefix. The model for calculating the conditioned probabilities is based on drawing without replacement. There is a bag with $n$ symbols of $k$ discriminable kinds. $n_a$ denotes how many symbols of kind $a$ are initially in the bag and $n'_a$ is the current number. The probability to draw a symbol of type $a$ is $n'_a/n$. If we pick a symbol $a$ both $n$ and $n'_a$ decrement by $1$. %\begin{equation} %\pmfn{\bar{A}}{n}(a^n) = \left\lbrace %\begin{array}{l l} %\frac{1}{\nTypePAcard} & \quad \text{if $a^n \in \nTypePA{n}{\pmf{\bar{A}}}$}\\ %0 & \quad \text{else} %\end{array} \right. %\end{equation} \begin{figure} \centering \input{diagrams/ccadm_refine} \caption{Refinement of the output intervals. Round brackets indicate symbols that must follow with probability one.} \label{fig:ccadm_refine} \end{figure} \begin{example}\label{ex:arithmRefine} Fig. \ref{fig:ccadm_refine} shows a refinement of the output intervals. Initially there are 2 '0's and 2 '1's in the bag.\The distribution of the first drawn symbol is$\pmfd{\bar{A}}{1}(0) =\pmfd{\bar{A}}{1}(1) = \frac{1}{2}$. When drawing a '0', there are 3 symbols remaining: one '0' and two '1's. Thus, the probability for a '0' reduces to 1/3 while the probability of '1' is 2/3. If two '0's were picked, two '1's must follow. This way we ensure that the encoder output is of the desired type. Observe that the probabilities of the next symbol conditioned on the previous symbols are unequal in general, i.e, we have \begin{equation} P_{\rv{\bar{A}}_2|\rv{\bar{A}}_1}(0|0) \neq P_{\rv{\bar{A}}_2|\rv{\bar{A}}_1}(0|1) \end{equation} in general. However, $\pmfn{\bar{A}}{n} = \prod_{i =1}^{n} P_{\rv{\bar{A}}_1|\rv{\bar{A}}^{i-1}}(a_i|a^{i-1})$ is constant on $\nTypePA{n}{\pmf{\bar{A}}}$ as we show in the following proposition. \end{example} \begin{proposition} After n refinements of the output interval the model used for the refinement step stated above creates equally spaced (equally probable) intervals that are labeled with all sequences in $\nTypePA{n}{\pmf{\bar{A}}}$. \end{proposition} \begin{proof} All symbols in the bag are chosen at some point. Consequently only sequences in $\nTypePA{n}{\pmf{\bar{A}}}$ may appear. All possibilities associated with the chosen string are products of fractions $n'_a/n$, where $n$ takes on all values from the initial value to $1$ because every symbol is drawn at some point. Thus for each string we obtain for its probability an expression that is independent of the realization itself: \begin{equation} \pmfn{\bar{A}}{n}(a^n) = \frac{n_{a=0}! \dotsb n_{a=k-1}!}{n!} = \frac{1}{\nTypePAcard} \quad \forall a^n \in \nTypePA{n}{\pmf{\bar{A}}}. \end{equation} \end{proof} Numerical problems for representing the input interval and the output interval occur after a certain number of input bits. For this reason we introduce a \emph{rescaling} each time a new output symbol is known. We explain this next. \subsection{Scaling input and output intervals} After we identify a prefix, we are no longer interested in code sequences that do not have that prefix. We scale the input and output interval such that the output interval is [0,1). Fig.~\ref{fig:ccadm_scaling} illustrates the mapping of intervals (in$_1$, out$_1$) to (in$_2$, out$_2$). The refinement for the second symbol works as described in Example \ref{ex:arithmRefine}. If the second input bit is $0$, we know that $10$ must be a prefix of the output. The resulting scaling is shown in Fig.~\ref{fig:ccadm_scaling} as (in$_2$, out$_2$) to (in$_3$, out$_3$). A more detailed explanation of scaling for arithmetic coding can be found for instance in \cite[Chap.~4]{sayood2006introduction}. We provide an implementation of ccdm online~\cite{website:ccdm}. \begin{figure} \centering \input{diagrams/scaling} \caption{Scaling of input and output intervals in case the input interval is a subset of an output interval. The latter interval corresponds to $[0,1)$ after scaling. A star indicates that this is just a prefix of the complete word. Round brackets indicate symbols that must follow with probability one.} \label{fig:ccadm_scaling} \end{figure} \section{Conclusion} We presented a practical and invertible f2f length distribution matcher that achieves the maximum rate asymptotically in the blocklength. In contrast to matchers proposed in the literature \cite{kschischang1993optimal,Ungerboeck2002,bocherer2011matching,amjad2013fixed,Cai2007, bocherer2013arithmetic} the f2f matcher is robust to synchronization and variable rate problems. Error propagation is limited by the blocklength. In future work we plan to investigate f2f length codes that perform well in the finite blocklength regime. \section{Acknowledgment} We wish to thank Irina Bocharova and Boris Kudryashov for encouraging us to work on the presented approach. %\section{Arithmetic Coding} %Arithmetic coding is a very general coding scheme that may be used for many different purposes. %It uses two probabilistic models one for input sequences and one for output sequences. %We restrict ourselves to probabilistic models where $\pmfn{A}{n}(a^n)$ can be expressed as $\pmfn{A}{n-1}(a^{n-1}) \cdot P_{\mathsf{A}_n|\mathsf{A}^{n-1}}(a_n)$ and either the input or input and output accept only fixed length codes. % %An arithmetic coder puts all possible input sequences (allowed by the input model) and all output sequences (allowed by the output model) vis-a-vis and adds an interval$[f,f+p)$ to each sequence, %with $p$ is the probability of the sequence and $f$ is the cumulative (lexicographically ordered, same length) probability of the last sequence. % %Two result of this allocation are that intervals of sequences that are a a prefix of other sequences include the intervals of latter sequences and %arithmetic coder order all intervals lexicographically. %All input intervals of same length string up to intervals that span from [0,1) each because all probabilities add up to 1. This holds for the output intervals of same length as well. % %The arithmetic coder finds for each input interval the biggest output interval that can be uniquely identified by the input interval. We will show possible methods in the examples and motivate differences for arithmetic coding % %\textred{This is the conceptual behavior of the arithmetic coder. %In practice the algorithm processes the input gradually, refines the input interval step by step and omits useless intervals. This way the algorithm avoids numerical problems. %A quick example. In Figure \ref{fig:ccadm_overview} we already know that if the input starts with 1 (0) the output starts with a (b). %We are in this case no longer interested in code sequences that do not have the same prefix. We scale the intervals such that the borders of the sure prefix now correspond to 0 and 1. In case the first input bit is 1, the lower bound of the interval labeled by '1001' will correspond to 0. %A detailed explanation can be found in \cite{bocherer2013arithmetic}} rewrite. % % % % %\begin{example}[Data Compression with iid source] %In data compression the output model is always iid \beronehalf distributed codewords. In every refinement step the probability of '0' or '1' is 0.5. %If the source is modeled to be iid then $ P_{\mathsf{A}_n|\mathsf{A}^{n-1}}(a) = \pmf{A}(a)$ for any input string $a^n$ and symbol $a$ in the input alphabet. %A output sequence is acceptable if its interval is a subinterval of the input intervals. If there is a stop symbol it is sufficient if the lower bound of the codesequence is element of the input interval. %\end{example} %\begin{example}[Data Compression on source with side information] %We can also use arithmetic coding when $ P_{\mathsf{A}_n|\mathsf{A}^{n-1}}(a) = \pmf{A}(a) $ is not true in general \textred{cite sayir} or when there is also side infomation available to both encoder and decoder. This kind of arithmetic coding is called adaptive arithmetic coding or context based arithmetic coding. The stream itself can be considered as side information. In data compression arithmetic coding is widely used because the data model for the input stream can be easily adapted \cite[Sec.~3.1]{Sayir1999coding}. %\end{example} %\begin{example}[arithmetic distribution matching] %in distribution matching the input bits are modeled iid \beronehalf. %As the aim of distribution matching is to transform the input to iid output symbols it is intuitive to choose an iid process for the output as well. %Although rather intuitive and simple (and therefore brilliant) this idea is pretty recent \cite{bocherer2013arithmetic}. However this results into a fixed-to-variable length code %\end{example} % % %\subsection{Constant Composion Code indexing} %We use arithmetic coding for indexing constant composition codwords efficiently. %This is done by using an adaptive model for the output. % %In case of distribution matching the statistics of the input are assumed to be \beronehalf. %We are interested in controlling the output sequence. %For this reason we connect a statistical model to the output that is depends on earlier output symbols. % %Remark, that we only consider a subset of all constant composition codewords $\nTypePA{n}{\pmf{\bar{A}}}$ of cardinality $2^m$ with $2^m \leq \nTypePAcard < 2^{m+1}$ according to \eqref{eq:bound_kc}. %It is not possible to index the whole set of constant composition codewords as there are not enough indeces unless $2^m = \nTypePAcard $. %The subset is unknown but distinctly defined by the arithmetic coder. %It is not necessary to determine the exact set as all elements of the whole set and therefore also all elements of the subset fulfill the constant composition constraint. % %The output model is based on drawing without replacement. %There is a bag with $n$ symbols of $k$ discriminable types. %$n_a$ denotes how many symbols of type $a$ are in the bag and $n'_a$ is the current number. %The probability to draw a symbol of type $a$ is $n'_a/n'$. %If we pick a symbol $a$ both $n'$ and $n'_a$ decrement by $1$. % %\begin{example}\label{ex:arithmRefine2} %Figure \ref{fig:ccadm_refine} shows refining. %Initially there are 2 '0's and 2 '1's in the bag. The probability for both symbols consequently is one half (first line diagram). %When drawing a '0', there are 3 symbols left, one '0' and two '1's. %Thus, probability for an '0' reduces to 1/3 while the probability of '1' is 2/3. %If two '0's were picked, two '1's must follow. %This way we ensure the coder generates only constant composition codwords %\end{example} % % %\begin{proposition} % The model used for the refinement step stated above creates equally spaced intervals that are labled with constant composition codewords. %\end{proposition} %\begin{proof} % All symbols in the bag are chosen at some point. % Consequently only constant composition codewords may appear. % All possibilities associated to the chosen string are always products of fractions $n'_a/n'$, with $n'$ takes all values from the initial value $n$ to $1$ because every symbol is drawn at some point. Thus for each string we obtain an expression that is independent of the realization itself % \begin{equation} % \pmfn{\tilde{A}}{n}(a^n) = \frac{n_0! \cdot \ldots \cdot n_{k-1}!}{n!} = \frac{1}{\nTypePAcard} \quad \forall a^n \in \nTypePA{n}{\pmf{\bar{A}}}. % \end{equation} %\end{proof} \bibliographystyle{IEEEtran} \bibliography{IEEEabrv,confs-jrnls,references} \end{document}}

56 — 1503.05146

\caption{Cable system of Sec.~\ref{sec:example1}: zero-sequence inductance (top panel), and resistance (bottom panel) obtained with MoM-SO ({\color{blue} ${\bf \mathlarger \circ}$}), FEM ($\cdot$), and analytic formulas ({\color{red} \xdash[.4em]~\xdash[.4em] }). Cable screens are left open.}

\caption{Cable System of Sec.~\ref{sec:example1}: node 1 and node 2 voltages obtained with MoM-SO ({\color{black} \xdash[.8em]}) and analytic formulas ({\color{red} \xdash[.4em]~\xdash[.4em] }) for the setup shown in Fig.~\ref{fig:3sccable_transient_setup}. }

\caption{Cable system of Sec.~\ref{sec:example2}: zero-sequence p.u.l. inductance (top panel) and resistance (bottom panel) obtained with the three-layer air-sea-seabed model in MoM-SO ({\color{blue} ${\bf \mathlarger \circ}$}), three-layer air-sea-seabed model in FEM ($\cdot$), two-layer air-sea model ({\color{red} \xdash[.4em]~\xdash[.4em] }), and two-layer sea-seabed model ({\color{magenta} \xdash[.4em]~\xdash[.1em]~\xdash[.4em]}). The screens are continuously grounded. }

\caption{Cable system of Sec.~\ref{sec:example2}: zero-sequence inductance (top panel), and resistance (bottom panel) obtained with the three-layer air-sea-seabed model in MoM-SO ({\color{blue} ${\bf \mathlarger \circ}$}), three-layer air-sea-seabed model in FEM ($\cdot$), two-layer air-sea model ({\color{red} \xdash[.4em]~\xdash[.4em] }), and two-layer sea-seabed model ({\color{magenta} \xdash[.4em]~\xdash[.1em]~\xdash[.4em]}). Cable screens are left open. }

\caption{Example of Sec.\ref{sec:example2}: voltages predicted at nodes 1 (top panel) and 2 (bottom panel) of the configuration in Fig.~\ref{fig:3sccable_transient_setup}. Plots compare the results obtained with three different ground models: three-layer air-sea-seabed({\color{black} \xdash[.8em]}), two-layer air-sea ({\color{red} \xdash[.4em]~\xdash[.4em] }), and two-layer sea-seabed model ({\color{blue} ${\bf \mathlarger \circ}$}). }

57 — 1503.05200

\caption{Distribution of $D_\text{off}/R_\lambda$ for the brightest LRGs, when the RMCG and BLRG are different. Errors are calculated by bootstrap. The red solid curve shows the best fit assuming a Rayleigh distribution. Black crosses represent the distribution of all \redmapper{} cluster members. Black plus signs represent the distribution of all \redmapper{} cluster members in clusters in which the BLRGs are not the central galaxies. The fact that the black crosses and black plus signs trace similar distributions suggests that clusters with non-central BLRGs are not dominated by projection effects and/or mergers. Grey curves represent the expected distribution of satellites distributed according to NFW profiles with different halo masses at redshift $z=0.25$. Grey solid, dashed, and dash-dotted curves are $\log(M_{200b})=14.0,14.5,15.0$, respectively (and the halo concentrations are $c=7.2,6.5,5.3$). The distribution of $D_\text{off}/R_\lambda$ for non central BLRGs is significantly different (truncated at the outskirts) compared to the distribution for all cluster members.}

58 — 1503.05768

\caption{Run time comparison for image denoising (in seconds) with different implementations. (1) The run time results with \colorbox{grayB}{gray}background are evaluated with the single-threaded implementation on Intel(R) Xeon(R) CPU E5-2680 v2 @ 2.80GHz; (2) the {\color{blue}{blue}} colored run times are obtained with multi-threaded computation using Matlab \textit{parfor} on the above CPUs; (3) the run time results colored in {\color{red}{red}} are executed on a NVIDIA GeForce GTX 780Ti GPU. We do not count the memory transfer time between CPU/GPU for the GPU implementation (if counted, the run time will nearly double)}

59 — 1503.07258

\caption{Pole locations of the nominal \textcolor{blue}{(blue)} and damaged aircraft \textcolor{red}{(red)}}

60 — 1503.08295

\caption{Multicomponent fits for 3C 57. \hb\and\mgii\from CAHA spectrum and\ciii\and\civ\from HST spectrum. The upper abscissa is rest-frame wavelength in\AA, the lower abscissa is in radial velocity units, and the ordinate is specific flux per unit wavelength in arbitrary units. The vertical long dashed line indicates the adopted rest frame. The black lines show the original continuum-subtracted spectra while the dashed magenta indicates the fit to the entire spectrum. The thick black and thin green lines show the broad and narrow components respectively. The blue-shifted component is indicated by a dashed blue line when detected. The light grey lines trace \feiiuv\and\feiiopt\emission which is considered in all four fittings. In some of them (as\hb\and\mgii) the contribution is very important but in others (as \ciii\and\civ) it is negligible. In panel (C) apart of \ciii\we show\siiii, \aliii\and\siii\in orange, dark green and yellow respectively. In panel (D) together with\civ\we plot the\niv\component in cyan and the\heiiuv\broad and blue component in brown and violet respectively.}

\caption{Object from \citet{sulenticetal07} with a \civ\blueshifts$\le$ -1000 \kms}

61 — 1503.09022

\caption{\label{fig:log_prob} With a linear base learner, and non-noisy data $\x_i \in \{0,1\}^2, \y_i \in \{0,1\}^3$, model \ref{fig:log0} and \ref{fig:log1} perform equally poorly (50\% under an \textit{exact match} measure), whereas \ref{fig:log2} performs perfectly in the same scenario. A human expert knows that there is no conditional dependence between \textsc{xor} and \textsc{or} (they are independent of each other given $\x$): this dependence is introduced by using an inadequate (linear) base learner, but even detecting or assuming this dependence will not necessarily lead to an improvement. {\pink It is worth noting that even \texttt{PCC} with Bayes-optimal search cannot solve \ref{fig:log0} and \ref{fig:log1}.}}

\caption{\label{tab:ccVxcc}\pink The average predictive performance of \texttt{BR}, \texttt{CC} and \texttt{CCASL} (with $H=L$) on 150 synthetic datasets of $L=10$ labels, $100$ hidden units (only in the case of the complex data) and $N=10000$ examples under 50/50 train/test splits.}

\caption{\label{tab:results_syn1} \pink Illustrative results on synthetic data of dimensions $D=2$, ($H=10$), $L=2$. The legend in the plots indicates $y_1y_2$. For the complex data generated with a hidden layer (above) we have provided exact match performance of \texttt{BR}, \texttt{CC} (with both possible chain orderings), and \texttt{CCASL}, and also the \textsf{acc}uracy of individual classifiers within \texttt{BR} and \texttt{CC}, which indicate marginal and conditional dependence. For the simple data (below) we already know that no dependence exists. Exact match performance between \texttt{BR} and \texttt{CC} methods (regardless of chain order) varies depending on the level of conditional dependence.}

\caption{\label{fig:dcc} \pink \texttt{CCASL} with an Augmented Middle Layer (\texttt{CCASL+AML}). }

\caption{\label{tab:methods}\pink Summary of methods. We propose the ones based on \texttt{CCASL}. }

\caption{\label{fig:contrast} \pink Different methods compared on the \textsf{Logical} dataset for varying numbers of hidden nodes. Each point represents the average of 10 runs on 60/40 train/test split with randomly shuffled labels.}

62 — 1504.00340

\caption{Correlations between the results coming from the three methods applied to Case Study 2. The three panels give the correlations between the outputs of (a) IBA and SA; (b) CWC and IBA; (c) SA and CWC. The coefficients of determination are given respectively by $R^2=0.5701$; 0.8807 and 0.3772. \red{Panel (d) is a Bland-Altman or Tukey mean-difference plot of differences between between results from pairs of approaches against their averages. The symbols ``$+$'' (red) compare CWC to IBA ($\mathcal{V}_{\rm{CWC}} - \mathcal{V}_{\rm{IBA}}$ vs $( \mathcal{V}_{\rm{IBA}} + \mathcal{V}_{\rm{CWC}})/2 $); ``$\times$'' (green) compare IBA to SA ($\mathcal{V}_{\rm{IBA}} - \mathcal{V}_{\rm{avg}}$ vs $( \mathcal{V}_{\rm{avg}} + \mathcal{V}_{\rm{IBA}})/2 $); ``$\circ$'' (blue) compare SA to CWC ($\mathcal{V}_{\rm{avg}} - \mathcal{V}_{\rm{CWC}}$ vs $( \mathcal{V}_{\rm{CWC}} + \mathcal{V}_{\rm{avg}})/2 $).} }

63 — 1504.00484

\caption{{ \color{NavyBlue} \bf Sketch of the experiment.} A quarter of the small annular cavity of width $D$ and radius $R$. The oil level is adjusted to obtain a depth $H$ in the cavity and a thin layer $H_0$ elsewhere. A walking droplet tends to remain in the cavity. Contours of the liquid surface $\zeta(\vec r, t)$ are shown for illustrating that the propagation of waves mostly takes place in the cavity while evanescent waves are observed outside the cavity. }

\caption{{ \color{NavyBlue} \bf String of walking droplets.} (a) Picture of a group of 7 droplets in the small annular cavity. One may observe quantified interdistances and antisynchroneous bounces for the successive droplets. (b) Red dots represent the droplet interdistances as a function of the label $n$ describing the interaction mode. The line is a fit using Eq.(\ref{eq:lambdaF}). Error bars are not indicated since they are smaller than the symbol size. Blue triangles are droplet interdistances given by the model for synchroneous (triangle up) or antisynchroneous bounces (triangle down). The dashed line is a fit using Eq.(\ref{eq:lambdaF}). }

\caption{{ \color{NavyBlue} \bf Speed for a pair of walkers.} (Dots) Speed of a droplet pair $v_2$ as a function of the distance $s$ between droplets. The speed is normalized by the speed $v_1$ of a single droplet. Error bars are indicated. (Triangles) The model explained in the main text returns quantified interdistances $s$ between droplets as well as specific speeds $v_2$ for both synchronous (triangle up) and antisynchronous (triangle down) cases. An excellent agreement is found between the model and the experimental data. The dashed curve is an exponential decay fitting the results from the model. }

\caption{{ \color{NavyBlue} \bf Speed for a string of walkers.} (Dots) Speed $v_N$ of a droplet string as a function of its number $N$ of components. The speed is normalized by the speed $v_1$ of single droplets. Error bars are indicated. The speed seems to saturate for large systems. (Triangles) The model described in the main text captures this effect for both synchronous (triangle up) and antisynchronous (triangle down) cases. An excellent agreement is found between the model and the experimental data. The dashed curve is a guide for the eye. }

64 — 1504.00722

\caption{$\mathcal{E}_A$ vs. time, $n=\{500,1000,5000\}$, Left : $k$ sparse, Right : $k$ dense, piecewise constant half-planes, \colorlegend}

\caption{$\mathcal{E}_A$ vs. time, $n=\{500,1000,5000\}$, Left : $k$ sparse, Right : $k$ dense, piecewise constant half-planes, \colorlegend}

\caption{$\mathcal{E}_A$ vs. time, $n=\{500,1000,5000\}$, Left : $k$ sparse, Right : $k$ dense, Gaussian density \colorlegend}

\caption{$\mathcal{E}_A$ vs. time, $n=\{500,1000,5000\}$, $k = 50, 150, 250, 450$, 3D half-cube density \colorlegend}

65 — 1504.00877

\caption{A full colour image of functions~\eqref{eq:=f+tf-t-}. The top picture shows \(F(z)\) followed by \(F^+(z)\) and \(F^-(z)\). We use a colour scheme developed by John Richardson. {\color{red} Red} is real, {\color{blue} blue} is positive imaginary, {\color{green} green} is negative imaginary, black is small magnitude and white is large magnitude. Branch cuts appear as colour discontinuities. Produced using MATLAB package \texttt{zviz.m}. }

66 — 1504.00964

\caption{~\\ Algorithm for efficiently computing $t^*$ in the general case. Comments are displayed in {\color{commentcol} gray}.}

67 — 1504.01108

\caption{Contour lines for the real and imaginary parts of function $F$~\eqref{eq:F-second-ex} and its rational approximation \(\tilde{F}\). They are superimposed on a full colour image using a colour scheme developed by John Richardson. {\color{red} Red} is real, {\color{blue} blue} is positive imaginary, {\color{green} green} is negative imaginary, black is small magnitude and white is large magnitude. Branch cuts appear as colour discontinuities and coalescent contour lines. Produced using MATLAB package \texttt{zviz.m}. }

68 — 1504.01122

\caption{{\cred{Experimental set-up for the compression testing of post-tensioned structures: L-prism (left), and 3LRL-column (right)}}}

69 — 1504.02239

\caption{\label{fig_STM}(Color online) Cross-sectional STM topography images of GaAs quantum dots capped with Al$_{0.3}$Ga$_{0.7}$As. The \textcolor{red}{white dotted} % %solid lines are guides to the eye that highlight the dot-barrier interface. }

\caption{\label{fig_Stark}(Color online) (a) Calculated Stark shift for a 12-nm-high GaAs quantum box surrounded by an infinite barrier (lines) %for $m^*_e=0.067$ and $m^*_h=0.5$ and a finite barrier (open circles) as a function of a vertical field. (b) Monte Carlo simulation for electric fields induced by randomly positioned surface charges with a density of $1 \times 10^{11}$~cm$^{-2}$ \textcolor{red}{at a point 60 nm} from the charge layer. The red line shows a field strength induced with a uniform charge sheet. }

\caption{\label{fig_th_broadening} (Color online) Calculated energy fluctuations due to randomly positioned surface charges with different densities of $1 \times 10^{11}, \, 5 \times 10^{11}$, and $1 \times 10^{12}$~cm$^{-2}$. The experimentally measured linewidths are also indicated by the gray points, \color{red}{which are equivalent to the data points shown in Fig.~\ref{fig_width_vs_e}}. }

70 — 1504.02338

\caption{Illustration of linear and kernel manifold alignment on the toy experiments. Left to right: data in the original domains (X1 = $\red{\bullet}$, X2 = $\blue{\bullet}$) and {\em per} class ($\red{\bullet}$, $\green{\bullet}$ and $\blue{\bullet}$), data projected with the linear and the RBF kernels, and error rates as a function of the extracted features when predicting data for the first (left inset) or the second (right inset) domain (\blue{KEMA$_{\text{Lin}}$}, \red{KEMA$_{\text{RBF}}$}, \cyan{SSMA}, \green{Baseline}).}

\caption{Example of the three first dimensions of the latent space in the \texttt{A} $ \to $ \texttt{W} experiment (top) and in the \texttt{C} $ \to $ \texttt{A} experiment (bottom). Left: by domain (\red{$\bullet$ = Source}, \blue{$\times$ = Target}), right: by class (each color represents a different class).}

71 — 1504.02340

\caption{Bounding box distance and appearance similarity are popularly used affinity metrics in the multiple target tracking literature. However, in real-world crowded scenes, they are often ambiguous to successfully distinguish adjacent or similar looking targets. Yet, the optical flow trajectories provide more reliable measure to compare different detections across time. Although individual trajectory may be inaccurate (\textcolor{red}{red} line), collectively they provide strong information to measure the affinity. We propose a novel Aggregated Local Flow Descriptor that exploits the optical flow reliably in the multiple target tracking problem. The figure is best shown in color.}

\caption{Our NOMT algorithm solves the global association problem at every time frame $t$ with a temporal window $\tau$. Solid circles represent associated targets, dashed circles represent unobserved detections, dashed lines show finalized target association before the temporal window, and solid lines represent the (active) association made in the current time frame. Due to the limited amount of observation, the tracking algorithm may produce an erroneous association at $t_2$. But once more observation is provided at $t_3$, our algorithm is capable of fixing the error made in $t_2$. In addition, our method automatically identifies new targets on the fly (\textcolor{red}{red} circles). The figure is best shown in color.}

\caption{Illustrative figure for unidirectional ALFDs $\rho'(d_i, d_j)$. In the top figure, we show detections as colored bounding boxes ($d_{\textcolor{red}{red}}$, $d_{\textcolor{blue}{blue}}$, and $d_{\textcolor{green}{green}}$). A pair of circles with connecting lines represent IPTs that are existing in both $t$ and $t + \triangle t$ and located inside of the $d_{\textcolor{red}{red}}$ at $t$. We draw the accurate (\textcolor{green}{green}), outlier (black), and erroneous (\textcolor{red}{red}) IPTs. In the bottom figure, we show two exemplar unidirectional ALFDs $\rho'$ for ($d_{\textcolor{red}{red}}$, $d_{\textcolor{blue}{blue}}$) and ($d_{\textcolor{red}{red}}$, $d_{\textcolor{green}{green}}$). The \textcolor{red}{red} grids ($2\times2$) represent the IPTs' location at $t$ relative to $d_{\textcolor{red}{red}}$. The \textcolor{blue}{blue} and \textcolor{green}{green} grids inside of each \textcolor{red}{red} bin ($2\times2 + 2$ external bins) shows the IPTs' location at $t+\triangle t$ relative to the corresponding boxes. IPTs in the grid bins with a \textcolor{red}{red} box are the one observed in the same relative location. Intuitively, the more IPTs are observed in the bins, the more likely the two detections belong to the same target. In contrast, wrong matches will have more supports in the outside bins. The illustration is shown using $2\times2$ grids to avoid clutter. We use $4\times4$ in practice. The figure is best shown in color.}

72 — 1504.02454

\caption{\blue{Bottom: Calibration data near the 2615~keV $^{208}$Tl $\gamma$-ray line, integrated over all bolometer-datasets. The solid blue line is the projection of the UEML fit described in the main text. In addition to the double-Gaussian lineshape for each bolometer-dataset, the fit function %also includes terms to model a multiscatter Compton continuum, a $\sim$\,30~keV Te X-ray escape peak, and a continuum background; these components, summed over all bolometer-datasets, are indicated by the blue dashed lines (a), (b), (c), and (d), respectively. Top: Normalized residuals of the data and the best-fit model.}}

\caption{ \blue{Bottom: The best-fit model from the UEML fit %analysis (solid blue line) overlaid on the %energy spectrum of \BBless~decay candidates in CUORE-0 (data points); %for simplicity of presentation the data are shown with Gaussian error bars. The peak at $\sim$2507~keV is attributed to $^{60}$Co; the dotted black line shows the continuum background component of the best-fit model. Top: The normalized residuals of the best-fit model and the binned data. The vertical dot-dashed black line indicates the position of $Q_{\beta\beta}$.}}

73 — 1504.02611

\caption{Example action rule for commands in \groove (\groove's rule colour coding: \blue{dashed blue} elements only exist on the left-hand side of rule (thus will be deleted), \green{bold green} elements on the right-hand side (thus will be generated), black ones persist)}

74 — 1504.03123

\caption{\textbf{The order-disorder transition of the vortex lattice in NbSe$_2$ at 4.2\,K.} The vortex lattice of NbSe$_2$, shown at 0.16, 0.18, 0.19, 0.20, 0.25, and 0.50\,T (cp. with figure~\ref{fig1}), becomes more disordered with increasing field. Each thin line indicates a bond between two adjacent vortices, the large (colored) symbols indicate lattice defects, namely vortices with four ({\color{violet} $\circ$}), five ({\color{red} $\circ$}), seven ({\color{blue} $\bullet$}), or eight ({\color{green} $\bullet$}) nearest neighbors. The horizontal bars correspond to a length of 1\,$\mu$m. The image boundaries were not used for the evaluation. \label{fig2}}

75 — 1504.03403

\caption{(Color online) The flow diagrams, Shields number $\Theta$ versus particle Reynolds number $Re_p$, with (a) $b=2$, (b) $b=4$, (c) and $b=6$. Diagonal lines of data points correspond to lines of constant $\Gamma$. The symbols show systems that came to rest (\textcolor{blue}{$\circ$}) or never stopped (\textcolor{green}{$\square$}) under protocol A, and were permanently (\textcolor{red}{$\times$}) or temporarily ($\bullet$) mobile as $\Theta$ was increased under protocol B. The dashed line shows $\Theta_c$, above which the inertial effects from particles entrained in the flow lead to sustained motion. The solid line indicates $\Theta_0$, below which the system will never be mobilized. The two large black open circles mark the parameter values we study in Fig.~\ref{fig:Weibull-scaling}.}

\caption{(Color online) Onset of bed motion is governed by Weibullian weakest-link statistics. (a)-(c) correspond to low $Re_p$ with $\Gamma=0.25$ and (d)-(f) correspond to high $Re_p$ with $\Gamma=0.1$. (a) and (d) show the probability distribution of the Shields number at bed failure $\Theta_f$ for many different system sizes ($W/D$ from 3.125 to 200, $N$ from 25 to 1600, and $WN/D$ from 8 to 80). The dashed vertical lines define (a) $\Theta_c$ and (d) $\Theta_0$. The insets show that $P(\Theta_f)$ collapses when rescaled by $\bar{\Theta}_f-\Theta_c$, where $\bar{\Theta}_f$ is the mean of each distribution. The solid line is a Weibull distribution with shape parameter $\alpha\approx 2.6$. (b) and (e) indicate that $\bar{\Theta}_f-\Theta_c$ for low $Re_p$ and $\bar{\Theta}_f-\Theta_0$ for high $Re_p$ scale with the effective system size $M_{\rm eff}^{-1/\alpha}$. (c) and (f) show that $t_m$, the mobilization time after bed failure, diverges near $\Theta_c$ and $\Theta_0$, respectively, independent of system size. The insets show a logarithmic plot of $t_m-t_{m,0}$ versus (c) $\Theta-\Theta_c$ and (f) $\Theta-\Theta_0$. The dashed lines show $t_m-t_{m,0} \propto (\Theta-\Theta_c)^{-0.43}$ (c) and $t_m-t_{m,0} \propto (\Theta-\Theta_c)^{-0.9}$ (f), and the thin vertical dashed line indicates $\Theta_c$. Symbols (\textcolor{yellow}{$\diamond$}, \textcolor{magenta}{$\star$}, \textcolor{cyan}{$\triangle$}, \textcolor{red}{$\circ$}, \textcolor{blue}{$\triangleleft$}, $\triangleright$, \textcolor{green}{$\square$}) correspond to different values of $N$ ($25$, $50$, $100$, $200$, $400$, $800$, and $1600$, respectively) with varying $W/D$. Each data point represents an average of 20 simulations.}

76 — 1504.03890

\caption{Each dot at position $(i,j)$ corresponds to a generally non-zero element at line $i$ and column $j$. The elements of the condition of order two $\{[D,a],[D,b^\circ]\}$ are black dots. The other dots describe the junk. The green dots correspond to $[D,a][D,b]c^\circ$, the pink dots to $[D,a^\circ][D,b^\circ]c$ and the blue dots to $ab^\circ$. Note that the \pink{non-zero} matrix elements of the junk and of the second-order condition do not \pink{overlap}. \label{figjunk}}

77 — 1504.03910

\caption{ \textcolor{blue}{Lorentz transformation of a relativistic hot plasma distribution. The bottom panel illustrates the flipping method, which is responsible for the spatial part of the Lorentz transformation.} \label{fig:shifted}}

\caption{ \textcolor{blue}{Distribution functions $f'({u}'_x)$ of Lorentz-boosted Maxwellians as a function of $u'_x$. Numerical results are overplotted on the analytic curves (Eq.~\ref{eq:f_ux}). We set $T=1$ for all cases.} \label{fig:f}}

78 — 1504.04313

\caption{The new charged states in the $c\bar{c}$ regions, ordered by mass. Masses $m$ and widths $\Gamma$ represent the weighted averages from the listed sources as in \cite{Brambilla:2014jmp}, or are taken from \cite{pdg} when available. The citation given in {\color{red} red} is for the first observation and the citation given in {\color{blue} blue} is for a non confirmation. The Status column NC (neds confirmation) indicates that the state has been observed by only one, or was not confirmed by other experiment. The Status is OK if at least two independent experiments saw the state.}

79 — 1504.04914

\caption{NCS-C( $T_{max}$, \boldsymbol{$\sigma$}, $r$, $epoch$, $N$)}

80 — 1504.06229

\caption{Different flow states for the forcing parameters $(f,A)$ scrutinised in \cite{dadamo2011b}. Visualisations are from \cite{thiria2006}. Solid lines represent the threshold from global to convective instability. The dotted line indicates the threshold to centrifugal instability of the Stokes layer of an oscillating cylinder without a crossflow given by $T_c=165$, from \cite{seminara1976}. Symbols \textcolor{verde}{$\bigstar$} show the threshold for 3D centrifugal instabilities observed in the 3D DNS discussed in Figs. \ref{scale_fig1-c} and \ref{scale_fig1-d}. Dashed region stands for states with turbulent-like behaviour described in \cite{dadamo2011b}. }

\caption{$|\phi|_{\max}$, red circle symbols, is the maximum of the Rayleigh discriminant modulus in function of the forcing phase $\alpha$, $|\phi|_{\max}=|\phi(\bar x_{\max} ,\alpha)|$. The lift coefficient $c_L$, green triangle symbols, gives a reference for the forcing phase $\alpha$. The local radius of curvature value $\mathcal R$, brown diamond symbols, at $\bar x_{\max}$ is close to the cylinder radius, its sign changes with the shedding cycle. Blue square symbols \textcolor{blue}{$\blacksquare$} mark three forcing phases depicted in Fig. \ref{dc2}.}

81 — 1504.08023

\caption{\textcolor{mod}{\textbf{Performance vs $\Delta$:} We plot performance on forecasting actions versus number of frames before the action starts. Our model (red) performs better when the time range is longer (left of plot). Note that, since our model takes days to train, we evaluate our model trained for one second, but evaluate on different time intervals. The baselines are trained for each time interval.}}

\caption{\textcolor{mod}{\textbf{Multiple Predictions:} Given an input frame (left), our model predicts multiple representations in the future that can each be classified into actions (middle). When the future is uncertain, each network can predict a different representation, allowing for multiple action forecasts. To obtain the most likely future action, we can marginalize the distributions from each network (right).}}

82 — 1504.08148

\caption{\emph{Selective state preparation procedure}. \textcolor{blue}{a)} A chain of $N$ closely spaced quantum emitters (separation $a$ with $k a \ll 1$, $k$ being the laser wave number) are individually driven with a set of pumps $\{\eta_j^m\}$. \textcolor{blue}{b)} The lasers are turned on for a time $T$, optimized such that an effective $\pi$-pulse into the desired subradiant target state is achieved. \textcolor{blue}{c)} Level structure for the $N$ systems where the $C_n^N$-fold degeneracy of a given $n$-excitation manifold is lifted by the dipole-dipole interactions. The target states are then reached by energy resolution (adjusting the laser frequency) and symmetry (choosing the proper $m$). \textcolor{blue}{d)} Scaling of the decay rates of energetically ordered collective states starting from the ground state (state index $1$) up to the single- and double-excitation manifolds for $6$ particles at a distance of $a=0.02\,\lambda_0$. The arrows identify the decay rates for the lowest energy states in the single (A) and double (B) excitation manifolds. \textcolor{blue}{e)} Numerical results of the time evolution of the target state population for $N=6$ and $a=0.02\,\lambda_0$ during and after the excitation pulse. Near unity population is achieved for both example states A (where we used $\eta=0.53\,\Gamma$) and B (for $\eta=2.44\,\Gamma$) followed by a subradiant evolution after the pulse time $T$ shown in contrast to the independent decay with a rate $\Gamma$ (dashed line).}

\caption{\emph{Coupling to dark states via a magnetic field gradient}. \textcolor{blue}{a)} Linearly increasing level shifts along the chain occuring in the presence of the magnetic field gradient. \textcolor{blue}{b)} Illustration of the level structure and indirect dark state access for two coupled emitters. While symmetry selects the state $\ket{S}$, off-resonant addressing combined with bright-dark state coupling of strength $\Delta_B$ allows for a near-unity population transfer into the state $\ket{A}$. \textcolor{blue}{c)} Dynamics in the single-excitation manifold of $N$ coupled emitters where symmetric driving reaches the bright states with amplitudes $\chi_m$ while the magnetic field couples neighboring dark and bright states. \textcolor{blue}{d)} Plot of the asymmetric state population for the two-atom case as a function of the increasing magnetic field (solid line) compared to the steady-state approximation (dashed line) at numerically optimized time $T= 16.19\,\Gamma^{-1}$, with parameters $\eta=\,\Gamma$ and $a=0.05\,\lambda_0$. \textcolor{blue}{e)} For a chain of $N=4$ emitters, a $91\%$-efficient $\pi$-pulse to the most robust state can be achieved as demonstrated in the population evolution plot. The separation is chosen to be $a=0.025\,\lambda_0$, while $\eta=40\,\Gamma$ and numerical optimization is employed to find $\Delta_B=0.98\,\Gamma$.}

\caption{\textit{Entanglement properties}. \textcolor{blue}{a)} Comparison of the numerically computed von Neumann entropy (empty circles) of the reduced density matrix of the chain minimized over the atom index and the analytical expression for the entropy of the Dicke state (green circles), both for excitations $n=1$ and $n=\lfloor N/2\rfloor$ as a function of the atom number $N$ at distance $a=0.1\lambda_0$. \textcolor{blue}{b)} Depth of entanglement of the subradiant four-atom state (blue dot) prepared by the magnetic field gradient scheme (see Fig. \ref{fig2}\textcolor{blue}{e}). It clearly lies above the $k=3$ boundary indicating four-atom entanglement. The $k$-atom entanglement boundaries of the target state population $P_t$ as a function of the ground state population $P_G$ have been computed for the corresponding target state of a four-atom chain at distance $a=0.025 \,\lambda_0$.}

83 — 1504.08158

\caption{(Color online) (a) The cumulative waiting time of simulations where $\rho_0=1.0$. Averages over 50 runs. The times when the end monomers of polymers of different lengths escape scale as $\tau \sim N_0^{1.29}$ (solid line). Reaction coordinates at instants when the inner part of the polymer exerts no force to the bead at the pore (\textcolor{plotpurple}{$\blacksquare$}): $s(\tau_2^*) \sim s^{1.0866}$ (dotted gray line). Reaction coordinate when number of beads inside the capsid $N = g_0$ (\textcolor{plotbrown}{$\blacktriangle$}) (b) The cumulative waiting time vs normalized reaction coordinate $s/N_0$ for $\rho_0 = 1.0$ and different $N_0$. Squares (\textcolor{plotpurple}{$\blacksquare$}) and triangles (\textcolor{plotbrown}{$\blacktriangle$}) correspond to those in (a). (c) Waiting times $t_w$ and the inverse of the measured force at the pore $1/f$ (arbitrary units) for $\rho_0 = 1.0$ and $\rho_0 = 1.5$ when $N_0 = 200$. $V$ is the capsid volume. (d) Waiting times $t_w$ vs. normalized reaction coordinate $s/N_0$ for $\rho_0 = 1.0$ and $N_0 = 50, 100,$ and $200$.}

84 — 1505.00015

\caption{Images of the flow for three precession angles for the model with $i=70^\circ$, $\Phi=110^\circ$, $\beta=10^\circ$ and $h/r=0.1$. The left hand side pictures blueshift and the right hand side pictures flux with the polarization vector overlaid, normalised to the maximum observed polarization degree. The three precession angles pictured, in units of cycles, are $\omega=0$, $0.3125$ and $0.625$ from top to bottom respectively. The full movies for these images can be found at {\footnotesize \color{blue} \underline{\smash{http://figshare.com/articles/Polarization\_modulation\_gifs/1351920}} \color{black}}. Indivdual movies can alternatively be found on YouTube: {\footnotesize \color{blue} \underline{\smash{www.youtube.com/watch?v=Q2CwOGKVC9U\&feature=youtu.be}} \color{black}} (blueshift) and {\footnotesize \color{blue}\underline{\smash{www.youtube.com/watch?v=E3kYAnS3pQI\&feature=youtu.be}} \color{black}} (flux and polarization).}

\caption{Images of the flow for three precession angles for the model with $i=30^\circ$, $\Phi=180^\circ$, $\beta=10^\circ$ and $h/r=0.1$. The left hand side pictures blueshift and the right hand side pictures flux with the polarization vector overlaid, normalised to the maximum observed polarization degree. The three precession angles pictured, in units of cycles, are $\omega=0$, $0.3125$ and $0.625$ from top to bottom respectively. Full movies corresponding to these images can be found at {\footnotesize \color{blue} \underline{\smash{http://figshare.com/articles/Polarization\_modulation\_gifs/1351920}} \color{black}}. Individual movies can alternatively be found on YouTube: {\footnotesize \color{blue} \underline{\smash{www.youtube.com/watch?v=TSe-{}-iXofu8\&feature=youtu.be}} \color{black}} (blueshift) and {\footnotesize \color{blue} \underline{\smash{www.youtube.com/watch?v=GjlIRfkor\_s\&feature=youtu.be}} \color{black}} (flux and polarization).}

85 — 1505.00066

\caption{Joint \textcolor{minorEd}{reasoning} for Pose Induction.}

86 — 1505.00353

\caption{(a) Alignment optimization where transformed leaf templates are deleted iteratively; (b) Tracking optimization where leaf candidates are transformed iteratively to align with the edge map ($80+$ iterations are used in this synthetic example). Numbers under images are the iteration number. \textcolor{yellow}{Yellow}/\textcolor{green}{green} dots are the estimated outer/inner leaf tips. \textcolor{red}{Red} contour is $\bm{W}(\bm{U};\bm{p})$. \textcolor{blue}{Blue} box encloses the edge points matching $\bm{W}(\bm{U};\bm{p})$. The number on a leaf is the leaf ID. Best viewed in color.}

87 — 1505.00431

\caption{Primitive cells of three crystal structures used in the ANNNI model. Atoms are colored according to their stacking position along the [111] direction. {\red{(a)}} depicts the primitive cell of the fcc structure with only one atomic site. The cubic cell is shown for a better imagination of the lattice. {\red{(b)}} shows the primitive cell of the hcp structure containing two non-equivalent atoms (two atomic sites). {\red{(c)}} represents the primitive cell of the dhcp structure with four non-equivalent atoms. }

88 — 1505.00502

\caption{\redtext Image plane for COSMOS J095930+023427 calculated by Lenstool. The critical line is shown in red. Each image position is shown as a cross with a label. The center of the mass distribution is shown in gray at the center}

\caption{\redtext{Image plane for SDSS J1320+1644 calculated by glafic. The blue line is the critical line. Image positions are shown as red triangles. The centers of the masses are shown as black crosses. }}

\caption{ {\redtext Image plane for SDSS J1430+4105 calculated by glafic. The blue line is the critical line. Image positions are shown as red triangles. The center of the mass density is shown as a black cross.}}

\caption{ {\redtext Image plane for J1000+0021 calculated by glafic. The blue line is the critical line. Image positions are shown as red triangles. The center of the mass density is shown as a black cross.}}

89 — 1505.00522

\caption{ (Color online) % Nucleation process of the high-Q skyrmion with $Q=2$. % Snapshots of the topological charge density $q(\mathbf{x})$, the energy density $\epsilon(\mathbf{x})$, the spin-component distribution $m_z(\mathbf{x})$, and its close up at sequential times. % The DMI constant $D=2$ mJ m$^{-2}$. The spin current density $j=3\times 10^{12}$ A m$^{-2}$. The external magnetic field $B_z=250$ mT. % In the simulation, each cell corresponds to one spin, and the cell size is $1.5$ nm $\times$ $1.5$ nm $\times$ $1$ nm. % In the spin distribution panels, each arrow stands for four spins, while it stands for one spin in the insets. % The nucleation process of the high-Q skyrmion with $Q=2$ is found to occur in two steps. First, it starts when a high-energy-density part is localized to a lattice-scale area, which possesses almost $Q=-1$. % A few spins rotate by large angles in this area, making the topological number of the area almost zero. The resultant spin texture has $Q=1$. % Second, a similar phenomenon occurs, yielding the high-Q skyrmion with $Q=2$ after the relaxation. % \blue{See Ref.~\onlinecite{SI} for Supplementary Movie 4.} }

90 — 1505.00524

\caption{Experimental results obtained from the full state tomography of the TE$_{01}$ mode, for qualitative analysis by rotating polarizer (\textcolor{blue}{Media˜1}). (a) Tomography matrix obtained by projection of the mode into six polarization and the corresponding six OAM states, measurement procedure (\textcolor{blue}{Media˜2}). (b) The qubit vector density matrix showing the high degree of non-separability of the investigated mode. }

91 — 1505.00687

\caption{\label{fig:track} Given the video about buses (the ``bus'' label are not utilized), we perform IDT on it. \textcolor{red}{red} points represents the SURF feature points, \textcolor{green}{green} represents the trajectories for the points. We reject the frames with small and large camera motions (top pairs). Given the selected frame, we find the bounding box containing most of the moving SURF points. We then perform tracking. The first and last frame of the track provide pair of patches for training CNN. }

\caption{\label{fig:3d} Surface normal estimation results on NYU dataset. For visualization, we use green for horizontal surface, blue for facing right and red for facing left, i.e., \textcolor{blue}{blue $\rightarrow$ X}; \textcolor{green}{green $\rightarrow$ Y}; \textcolor{red}{red $\rightarrow$ Z.}}

92 — 1505.01085

\caption{\small Sample labelings from our automatic method. The action and point of contact is indicated with a red dot (e.g., upper right is the contact location of the pelvis when sitting upright). \textcolor{ForestGreen}{Green: Supports Affordance}; \textcolor{red}{Red: does not}; \textcolor{blue}{Blue: unknown or missing data.} Note that our method captures the {\it physical} affordances of the scene, so it permits standing on the counter and on the stools in the middle example, even though we would not expect a human to do so. }

\caption{Sample results from our mid-level patch model (\textcolor{red}{red: supports affordance}; \textcolor{blue}{blue:does not}). Our method captures distinctions between the affordances: for instance, one probably cannot touch anything standing on top of the desk in row 1. }

\caption{Example results from our CNN (\textcolor{red}{red: supports affordance}; \textcolor{blue}{blue: does not}). We infer a good interpretation of affordances. Notice areas where some affordances are possible but not others (e.g., stand vs. lie in row 1, under the desk).}

93 — 1505.01292

\caption{(Color online) Rietveld refinement (solid red line) of room temperature x-ray powder diffraction data for NpIr (\textcolor{blue}{$\circ$}) annealed at $873$~K. The difference between the calculated and experimental points is shown by the black line which has been offset by -2000~cts/4s for clarity. The vertical tick marks correspond to the Bragg peak positions for the NpIr $P2_1$ structure shown in the inset.}

\caption{(Color online) (a) Electrical resistivity of NpIr, with inset showing fits to a Fermi liquid (solid red line) and non-Fermi liquid (dashed green line) model at low temperatures. (b) Isofield magnetoresistance of NpIr at $14$~T, with inset showing the field dependence of the isothermal magnetoresistance for $T=1.8$~K (\textcolor{red}{o}), $3$~K (\textcolor{dkgreen}{$\Box$}), $5$~K (\textcolor{blue}{$\triangle$}), and $50$~K ($\star$). (c) Hall Effect measurements of NpIr, with inset showing fit to $R_H(T)=R_0+R_1\chi^*(T)$.}

94 — 1505.02012

\caption{Cooling of the particle motion: Closeup of the PSD in the region of the first harmonic at 40kHz of a particle at $2\times 10^{-5}\,\rm mBar$ without ( \crule[red!100!white!100]{0.25cm}{0.25cm} ) and with feedback ( \textcolor{cyan}{$\blacktriangledown$} ). Darker lines are Lorentzian fits. }

95 — 1505.02115

\caption{Momentum distribution functions of the conduction and $f$ electrons in the noninteracting case ($U_f=U_{cf}=0$, $V/W=0.1$, $\varepsilon_f=0$). The solid and dotted lines are obtained from Eq. (\ref{eq:nfk_Uf0}) and (\ref{eq:ndk_Uf0}), respectively. The symbols \textbullet, \textcolor{red}{$\blacksquare$} denote the DMRG results for $L=50$. }

\caption{Momentum distribution functions of the conduction (\textbullet) and $f$ (\textcolor{red}{$\blacksquare$}) electrons for $L=50$ and $n=2$. Panel (a), (b), (c) and (d) correspond to $U_{cf}/W=0,1.5,1.7$ and 3, furthermore $U_f/W=3$, $V/W=0.1$ and $\varepsilon_f=-U_f/2$ in all cases. The lines are guides to the eye.}

\caption{The $f$-level occupancy as a function of $\varepsilon_f$ for $L=80$, $U_{cf}=0$ (\textcolor{red}{$\blacksquare$}) and $U_{cf}/W=4$ (\textbullet), furthermore $U_f/W=10$, $V/W=0.2$. The circled data points are used in the comparison in Fig. \ref{fig:mom_dist_n1p75}. The lines are guides to the eye.}

\caption{Momentum distribution functions of the conduction (\textbullet) and $f$ (\textcolor{red}{$\blacksquare$}) electrons for $L=80$ and $n=1.75$. Panel (a), (b) correspond to $U_{cf}/W=0$, $\varepsilon_f/W=0.5$ and $U_{cf}/W=4$, $\varepsilon_f/W=-0.75$, respectively, furthermore $U_f/W=10$, $V/W=0.2$ in all cases. The lines are guides to the eye. }

96 — 1505.02128

\caption{Growth functions, $\sigma_{cV} t$ (\solidrule) and $\sigma_{mV}t$ ($\bigcirc$), for $\kappa' = 0, R = 2, w = 512, k = 0.15$. The solid dot ($\CIRCLE$) represents the onset of the nonlinearity. Inset: Magnified near the onset of instability. }

97 — 1505.02836

\caption{(Color online) \label{fig:bs_and_shifts}(a) Photonic bandstructure (in units of wave vector versus vacuum wavelength) of the ideal W1 showing the fundamental (solid) and higher order (dashed) guided modes. The light line is shown in black. (b) The broadened (blueshifted on average) bandstructure of the unperturbed fundamental mode (blue/dark-solid line) for extrinsic disorder of $0.02a(\SI{4.8}{\nano\metre})$. The three greyscale shades indicate the statistical distribution of the perturbed eigenfrequencies that lie within $\pm \sigma,\, \pm2\sigma,\, \pm3\sigma$ (dark-grey, medium-grey, light-grey) of the unperturbed fundamental mode, where $\sigma$ denotes the standard deviation. (c) Ensemble averaged first-order eigenvalue corrections for six disordered samples representing the mean net frequency shift. (d) Ensemble averaged second-order eigenvalue corrections representing the standard deviation of the mean net frequency shift. In both graphs, the \emph{intrinsic} disorder is kept fixed at \textcolor{cyan}{$0.005a(\SI{1.2}{\nano\metre})$} while the external disorder is varied as follows: \textcolor{cyan}{$0a$} (cyan/solid-light-grey), \textcolor{green}{$0.01a(\SI{2.4}{\nano\metre})$} (green/dashed-light-grey), \textcolor{red}{$0.02a(\SI{4.8}{\nano\metre})$} (red/solid-medium-grey), \textcolor{magenta}{$0.03a(\SI{7.2}{\nano\metre})$} (magenta/dashed-medium-grey), \textcolor{blue}{$0.04a(\SI{9.6}{\nano\metre})$} (blue/solid-dark-grey), \textcolor{black}{$0.05a(\SI{12}{\nano\metre})$} (dashed-black). % $0,\, 0.01a=2.4\,\mathrm{nm},\, % 0.02a=4.8\,\mathrm{nm},\, 0.03a=7.2\,\mathrm{nm},\, % 0.04a=9.6\,\mathrm{nm},\, 0.05a=12\,\mathrm{nm}$. }

98 — 1505.03076

\caption{\coloronlineornothing Sketch of the model system. A rigid object interacts with the substrate though the formation and breaking of bonds. Each bond can be in a broken, a strong or a number of weak configurations.\label{fig:sketch}}

\caption{\coloronlineornothing Example time evolution of object speed and bond properties when \fpf{a} only weak (but fast) bonds are possible, and when \fpf{b} both weak (but fast) and strong (but slow) bonds are possible. Top panels: object speed and average force in bonds. Bottom panels: fractions $\nu$ of unformed (\blackdash), weak (\cyandash) and strong (\reddash) bonds. In this paper $N=300$, $\eta=N/4$, $M=N$, $\kappa=[10,\, 10]$ (for $[\text{weak},\,\text{strong}]$ bonds), $l_0=h=1$, $\lambda=0.4\eta/M$, $\koffzero=[0.2,\,0.001]$, $\konzero=[0.1,\,0.1]$, $V_0=[\Vweakzero,\,\Vstrongzero]=[1,\,0.1]$, $a=[1,\,1]$, $\Delta\tau=[0.1,\,0.1]$, $\beta=1$, $\Delta x=[1,\,1]$, $\Delta t=0.01$. In this figure $\Ntype=[5,\,0]$ \fpl{a} or $\Ntype=[5,\,1]$ \fpl{b}, $\Vfluid=0.8$.\label{fig:time_evolution}}

\caption{\coloronlineornothing Average steady-state speed of rigid object vs fluid flow speed $\Vfluid$ for systems with different bonds available. \fpf{a} Strong only, weak only, and a combination. \fpf{b} Weak only and the corresponding weak plus one strong bonds. The gray dashed line shows $\Vobject=\Vfluid$ as a guide to the eye. Legends show values of $\Ntype$. $\Vobjectaverage$ was measured over the second half of simulations lasting for $150$ units of time. Points near the transition from arrest to non-arrest are averages over ca. 20 simulations.\label{fig:Vslider_vs_Vfluid}}

\caption{\coloronlineornothing Estimated object arrest time. \fpf{a--d} Distributions of arrest times for selected simulation settings (arrows originate on the $\Ntype=[3,\,1]$ and $[5,\,1]$ lines). The fits in \fpl{a} and \fpl{d} are normal distributions. Each histogram is based on ca. $2000$ simulations. \fpf{e} Estimated arrest time vs $\Vfluid$. Colors and markers are the same as in \fig~\ref{fig:Vslider_vs_Vfluid}. For the points close to where each line diverges, the total simulation time was increased to $300$ units of time. Error bars show one standard deviation away from the mean. Away from $\Vfluidtransition$ the error bars are smaller than the markers. Each point is based on 40 simulations.\label{fig:arrest_time}}

\caption{Average bond fractions reached in steady state. In the simulations with only weak bonds, the average fraction of formed bonds has an appreciable negative slope throughout the range where bonds are formed. In the simulations with strong bonds as well, the fraction of bonds formed stays nearly constant up to $\Vfluidtransition$, and then drops sharply. This is consistent with our observation in the main text that strong bonds arrest the object, but weak bonds do not, because the bond fraction depends on the slider speed, which becomes independent of $\Vfluid$ as long arrest occurs. Legend: fractions of unformed (\blackdash), weak (\cyandash) and strong (\reddash) bonds for $\Ntype=[1,\,0]$ (\diamondo), $[1,\,1]$ (\diamondf), $[3,\,0]$ (\circleo), $[3,\,1]$ (\circlef), $[5,\,0]$ (\squareo) and $[5,\,1]$ (\squaref).\label{fig:steady_state_bond_fractions}}

99 — 1505.03230

\caption{(Color online) \red{An illustration of the extracted} extreme points at different $s$ levels: a) $s=100$, and b) $s=200$. The local maxima and minima are demonstrated respectively by $\ocircle$ and $\square$. The vertical line indicates the window size $s$. \red{To ensure spacial homogeneity, a sliding window with level $s$ scans over the whole data set, see more detail in the text.}}

100 — 1505.03510

\caption{(color online) \red{The momentum distribution $a_p/a_0$ given in Eq.~(\ref{ap}) as a function of $pa_0$ for various values of $Z_0$. For $Z_0 < 1$ there is a negative correction, i.e. the continuum states try to make the approximation based on the bound states more extended, while for $Z_0 > 1$ the corrections are all positive, so try to increase the value of the wave function near the origin. The scale on which the continuum momentum eigenstates contribute most increases with increasing $Z_0$.}}

\caption{(color online) Various contributions from the Hydrogenic bound and continuum states to the ground state wave function for a central charge with magnitude $eZ_0$, as a function of $Z_0$. At $Z_0 = 1$ only the $1s$ state contributes, as expected. For $Z_0 \ne 0$ all other states contribute as well. In particular, for $Z_0 > 1$ the continuum states contribute with increasing amplitude for reasons explained in the text. The sum of the three curves is unity for all $Z_0$, \red{as indicated by the thick horizontal (purple) line across the top.}}

\caption{(color online) \red{The evolution of the ground state wave function from just considering the bound states, i.e. the lowest (blue) curve to the exact result (blue squares). As the accuracy increases we cut off the momentum integration in Eq.~(\ref{corr}) at $p_ca_0$ = 1 (next highest (red) curve), $p_ca_0$ = 3 (next highest (green) curve), $p_ca_0$ = 10 (next highest (black) curve), and finally, $p_ca_0$ = 50 (next highest (red) curve). The higher momentum continuum components contribute to the wave function near the origin.}}

\caption{(color online) \red{The ground state energy for a $Z_0{\rm Hydrogen}$ system with $Z_0 = 3$, vs. $Z_{\rm ref}$. This calculation was performed by diagonalizing a $18 \times 18$ matrix, and shows that using larger values of $Z_{\rm ref}$ is effective to a degree. Results are definitely more accurate than using $Z_{\rm ref} = 1$ for example, but the convergence to very accurate results is very slow.}}

101 — 1505.03949

\caption{\label{f:pair}The 2MASX occulting galaxy pair at z=0.06. The extinction in the overlap region ({\bf black aperture}) can be estimated from the complementary apertures; the foreground spiral ({\color{ForestGreen}F, the green aperture}) and the background galaxy ({\color{red}B, red aperture}).}

\caption{The relative difference between $\Lambda$CDM predicted distance and the SN\,Ia luminosity distances from SDSS-SN. Individual SN\,Ia (grey points) and the mean and rms (red points) and once corrected for$A_V$ bias (green points) using the occulting galaxy templates (Figure \ref{f:pav}), matched through host mass. Alternatively, one can match OCG templates by radius (blue points). Two Cosmologies are shown as well to illustrate: w=-0.9 and -1, neither a fit to the data. The different Cosmologies are shown to illustrate how the choice of dust prior can have an almost similar effect on the inferred Cosmology.}

102 — 1505.04143

\caption{Comparison of sparse keypoint localization for our method, SIFT Flow~\cite{liu_PAMI_2011} and Deep Flow~\cite{weinzaepfel_ICCV_2013}. The baseline measures the global alignment bias of the dataset (how well one would perform by simply assuming no flow). The argmax considers taking the single best match without regularization. The graphs measure the fraction of correspondences which fall within an increasing distance from groundtruth. $3$ standard deviations is inperceptible from human annotator accuracy. From left to right: (a) aggregate results across all images, (b) \textcolor[rgb]{0.157,0.647,0.827}{the truck pair} which our method localizes well, and \mbox{(c) \textcolor[rgb]{0.714,0.2,0.373}{the biking pair}} for which our method fails to produce any meaningful correspondences. \vspace{2mm} \label{fig:results} }

103 — 1505.04486

\caption{\label{fig:ss} (Color online) Stress-strain response of the simulated (Na$_2$O)$_{20}$(SiO$_2$)$_{80}$ glass during fracture, \textcolor{red}{compared with that of pure silica}. The snapshots illustrate the volume of the cavities, with a radius larger than 5 \AA, that are formed at various strains. The inset shows the relative variation of the average Si--O and Na--O bond lengths ($\Delta l/l_0$) with respect to the strain. The cutoffs used to identify Si--O and Na--O bonds are 1.974 \AA\and 3.311\AA, respectively, as determined from the position of the first minimum after the peak associated to the first coordination shell in the pair distribution functions. }

\caption{\label{fig:map} (Color online) Contour maps of the local number of constraints per atom ($n_{\rm c}$) over a 8 \AA-thick slab inside the 20NS structure, \textcolor{red}{before (left, $\epsilon=0$) and after (right, $\epsilon=0.4$) fracture. $n_{\rm c}$ is calculated from the local sodium oxide concentration on a square grid of 8 \AA\in resolution.} The grey area indicates the extent of the final crack. The plot on the upper left corner shows the standard deviation of $n_{\rm c}$ as a function of the \textcolor{red}{grid resolution}. }

\caption{\label{fig:plastic} (Color online) Probabilities of plastic events (see text) occurring around the bridging (BO, \textcolor{red}{open circles}) and non-bridging (NBO, \textcolor{red}{open squares}) oxygen atoms with respect to the local number of constraints per atom $n_{\rm c}$. The sampling frequency in terms of $n_{\rm c}$ is chosen so that each group contains at least 500 oxygen atoms. The grey area indicates the boundaries of the intermediate phase observed experimentally through modulated \textcolor{red}{Differential Scanning Calorimetry} \cite{Vaills2005, Micoulaut2008}. }

104 — 1505.05269

\caption{Comparative analysis of different sensor nodes\label{tab:four}}{\includegraphics[scale=.06]{table7-eps-converted-to}}

\caption{Merits and Demerits of different architecture design\label{tab:table5-eps-converted-to}}{\includegraphics[scale=.075]{table5-eps-converted-to}}

105 — 1505.05284

\caption{The values of $\err_\image^2$ and $\err_\chi$ are displayed in relation to the number of degrees of freedom in a log-log plot for the applications (a) (upper left), (b) (upper right), (c) (lower left) and (d) (lower right). The plotted error estimator values correspond to the discretizations (FE) ($\err_\image^2$ (black line, {\color{black} $\blacksquare$}) and $\err_\chi$ (brown line, {\color{gnuplotbrown} $ \bullet$})) and (FE') ($\err_\image^2$ (red line, {\color{gnuplotred} $\filledmedtriangleup$}) and $\err_\chi$ (blue line, {\color{gnuplotblue} $ \filledmedtriangledown$})), respectively.}

\caption{First row: Input image $u_0$ composed by the superposition of two Gaussian kernels, numerical solutions $\image_h$, $(p_h)_1$ and $(p_h)_2$ computed via the adaptive algorithm using discretization (FE'). Second row: $\chi_h$, deciles and the error estimators in a log-log plot ($\err_\image^2$ in red {\color{red} $\filledmedtriangleup$}, $\err_\chi$ in blue {\color{blue} $ \filledmedtriangledown$} for the discretization (FE') and $\err_\image^2$ in black {\color{black} $\blacksquare$}, $\err_\chi$ in brown {\color{gnuplotbrown} $ \bullet$} for the discretization (FE)).}

106 — 1505.05402

\caption{Lin-log plot of the velocity profiles in wall units, for shear Reynolds number $Re_{\tau}=185$ (left panel) and Rayleigh numbers $Ra=0$ ($\ast$), $Ra=8.12 \times 10^5$ (\textcolor{blue}{$\bullet$}) and $Ra=6.5 \times 10^6$ (\textcolor{red}{$\blacksquare$}). The dashed lines indicate the linear law, valid in the viscous layer, $U^+ = y^+$, while the dash-dotted ones correspond to the prediction (\ref{eq:tLog-Law})-(\ref{eq:Kc}) with $\kappa = 0.42$, $C_S=2.5$ and $B \in [5.8, 6.7]$ ($Pr=1$).}

107 — 1505.05850

\caption{\textbf{Buildup of correlated backaction noise during optical coupling.} \textbf{a} Distributions of about $ 10^4$ quadrature measurements on the two oscillators under three conditions: no coherent excitation or coupling (left) or coherent excitation and coupling of duration $\tau_c$ (center and right). Dashed lines indicate 1-$\sigma$ variance ellipses for signal (red) and optical shot noise (black); the excess signal variance quantifies the incoherent excitation. Optical coupling increases the incoherent excitation level, plotted in $\textbf{b}$ as the number of added phonons for the high- ({\color{blue}$\blacktriangle$}) and low-frequency ({\color{darkgreen}$\blacktriangledown$}) oscillator. The measured data show agreement to zero-free-parameter theory calculated for each $\tau_c$ (open symbols); see Supplementary Information. The coupling leads to correlations in the displacements of the two oscillators, quantified by the covariance $r_{}$ ({\color{red}$\blacklozenge$}) between their quadratures. Correlations show the expected trend, slightly reduced below zero-free-parameter expectations (open symbols). The measured 1-$\sigma$ correlations for uncoupled oscillators are indicated by grey shaded range. Solid lines indicate theory for average experimental parameters. Bars show typical 1-$\sigma$ standard statistical errors; a plot with all error bars can be found in the Supplementary Information.}

108 — 1505.06137

\caption{ (a) Shear stress response $\sigma (t)$ for different applied shear rates [color, $\dot \gamma$ (s$^{-1}$)]: [\textcolor{yellow}{\textbf{\large --}}, 100]; [\textcolor{orange}{\textbf{\large --}}, 150]; [\textcolor{red}{\textbf{\large --}}, 200]; [\textcolor{black}{\textbf{\large --}}, 250]. (b)-(e) Velocity profile $v(r)$ across the gap, where $r$ is the distance to the rotor at different times [symbol, time (s)]: [$\bigcirc$, 200]; [$\Box$, 400]; [$\bigtriangleup$, 800]; [$\bigtriangledown$, 1400]. Each color corresponds to each applied shear rate in (a). The rotor velocity corresponds to the upper bound of the vertical axis. The sample is left to age during $t_w = 60$~min before each experiment. (f) Shear stress response $\sigma (t)$ for different waiting times $t_w$ [color, $t_w$ (min)]: [\textcolor{yellow}{\textbf{\large --}}, 100]; [\textcolor{orange}{\textbf{\large --}}, 30]; [\textcolor{red}{\textbf{\large --}}, 5]; [\textcolor{black}{\textbf{\large --}}, 1]. Experiments performed at $\dot\gamma=$~100~s$^{-1}$. (g)-(j) Velocity profile at different times [symbol, time (s)]: [$\bigcirc$, 30]; [$\Box$, 400]; [$\bigtriangleup$, 800]; [$\bigtriangledown$, 1100]. Each color corresponds to each value of $t_w$ in (f). The rotor velocity corresponds to the upper bound of the vertical axis. }

\caption{Flow state diagram in the (sample age $t_w$, applied shear rate $\dot\gamma$) plane. In steady state, the gel may either be fully fluidized (\textcolor{blue}{$\blacklozenge$}, blue region) or display shear banding. Steady shear banding is represented in (\textcolor{red}{$\bullet$}) symbol which size encodes the portion of the gap that is being sheared. Unsteady banding, which denotes flows where the band width and the slip velocity at the rotor display significant fluctuations, is represented by (\textcolor{red}{$\blacklozenge$}). }

\caption{Full fluidization observed for $\dot\gamma=150$~s$^{-1}$ and $t_w=30$~min. (a) Shear stress $\sigma$ (\textcolor{black}{\textbf{\large --}}) and slip velocity $v_s$ (\textcolor{red}{\textbf{\large --}}~) vs time. (b)~Width $\delta$ of the fluidized shear band normalized by the gap width $e$ vs time. (c)~Local shear rate within the shear band (\textcolor{black}{\textbf{\large --}}) and global shear rate (\textcolor{red}{\textbf{\large --}}) \break vs time. The horizontal dotted line indicates the shear rate applied by the rheometer. In (a)-(c), the vertical dashed line indicates the fluidization time $\tau_f$.}

\caption{Unsteady shear banding observed for $\dot\gamma=50$~s$^{-1}$ and $t_w=30$~min. (a) Shear stress $\sigma$ (\textcolor{black}{\textbf{\large --}}) and slip velocity $v_s$ (\textcolor{red}{\textbf{\large --}}) vs time. (b)~Width $\delta$ of the fluidized shear band normalized by the gap width $e$ vs time. (c)~Local shear rate within the shear band (\textcolor{black}{\textbf{\large --}}) and global shear rate (\textcolor{red}{\textbf{\large --}}) vs time. The horizontal dotted line indicates the shear rate applied by the rheometer.}

\caption{(a)~Elastic ($\circ$) and viscous moduli (\textcolor{red}{$\bigtriangleup$}) vs time after a preshear at $\dot \gamma_p=500$~s$^{-1}$ for 2~min ($f=1$~Hz, $\sigma=0.05$~Pa). Same data as in Fig.~2(a) in the main text. (b)~Flow curve, shear stress $\sigma$ $vs.$ applied shear rate $\dot\gamma$, obtained by decreasing ($\bigtriangledown$) and increasing ($\bigtriangleup$) shear rate in order from 10$^{-2}$ to 10$^{3}$ s$^{-1}$ with a waiting time of $8$~s per point. Black (resp. red) symbols correspond to results in a smooth (resp. rough) geometry.}

\caption{Elastic ($\bigtriangleup$) and viscous ($\bigtriangledown$) moduli vs frequency $f$. The frequency sweep is performed at a constant strain $\gamma=0.5$~\% and at different aging times $t_w$ after preshear. [color, $t_w$ (min)]: (\textcolor{black}{\Large{-}},~30); (\textcolor{red}{\Large{-}},~60); (\textcolor{yellow}{\Large{-}}, 100). }

\caption{Elastic ($\circ$) and viscous ($\bigtriangleup $) moduli vs the strain amplitude. The strain sweep is performed at a fixed frequency $f=1$~Hz with a waiting time of 8~s per point, and at different times $t_w$ after the preshear. [color, $t_w$ (min)]: (\textcolor{black}{\Large{-}},~30); (\textcolor{red}{\Large{-}},~60); (\textcolor{yellow}{\Large{-}}, 100). The vertical dashed line at $\gamma \sim 7$~\% emphasizes the yield point defined as the point at which $G'=G''$.}

\caption{(a) Stress response $\sigma(t)$ to a shear startup experiment performed at $\dot \gamma=50$~s$^{-1}$ for different gap sizes: [color, gap (mm)] = [\textcolor{brown}{\textbf{\large --}}, 2], [\textcolor{red}{\textbf{\large --}},~1], [\textcolor{black}{\textbf{\large --}}, 0.5]. (b)-(d) Velocity profile, where $r$ is the distance to the rotor at different times in (a) [symbol, time (s)]: [$\bigcirc$, 100]; [$\Box$, 800]; [$\bigtriangleup$, 1200]; [$\bigtriangledown$, 2000]. The rotor velocity corresponds to the upper bound of the vertical axis. The sample is aged during $t_w=10$~min before each experiment. }

109 — 1505.06401

\caption{(Color online) Experimental dispersion of the energy-loss peak positions of the A$_1$ and B$_1$ excitons as function of momentum transfer in the $\Gamma$M (\textcolor{blue}{$\times$}) and $\Gamma$K (\textcolor{red}{$\medcircle$}) directions and fitted dispersion in the $\Gamma$M (\textcolor{blue}{\hdashrule[0.5ex]{0.8cm}{0.7pt}{0.4mm}}) and $\Gamma$K (\textcolor{red}{\hdashrule[0.5ex]{0.8cm}{0.7pt}{0.9mm}}) directions. The peak positions were obtained by applying polynomial fits to the data presented in Fig. \ref{fig4} and validated quantitatively.}

110 — 1505.06436

\caption{{\it Upper panel:} FOF mass functions for the \coco{} and the \scolor{} simulations, on which we superimpose the Sheth-Tormen prediction (solid lines). The \coco{} results and the ST predictions are plotted for a wide range of redshifts, from $z=9$ (red diamonds) to $z=0$ (purple circles). For comparison we also plot the $z=0$ results from the \scolor{} run (open hexagons) and the Reed {\it et al.} prediction (dashed line). The vertical bars indicate Poisson errors. {\it Bottom panel:} The \coco{} and \scolor{} mass functions normalized by the ST prediction at that redshift. The horizontal dashed lines mark a $10\%$ difference level. The vertical dashed (dotted) line illustrate the \coco{} (\scolor{}) halo mass resolution limit. }

\caption{{\it Left-hand panel:} Concentration-mass relation for relaxed haloes. The open symbols give the median concentration obtained from fitting NFW and Einasto profiles. The best-fitting line was obtained from the combined Einasto \coco{}+\scolor{} data, with the \coco{} concentrations (circles) extended by the \scolor{} results (triangles) for halo masses $M_{200}>5\times 10^{11}\Msun$. The filled circles show the average concentrations obtained from stacking profiles for the three lowest mass bins. For comparison we also show the data obtain by fitting the NFW profile (open boxes). {\it Right-hand panel:} Comparison of various $c_{200}-M_{200}$ fits from the literature (lines with symbols and the dashed line) with our best-fitting line (solid line). In both panels, the shaded region shows the 16th to 84th percentile scatter of the \coco{}+\scolor{} data. }

\caption{A comparison of the $V_{max}-M_{sub}$ relation for subhaloes found in hosts of all masses at redshift, $z=0$. The \coco{}, \scolor{} and \MII{} lines depict the median relation found by binning the subhaloes according to their $V_{max}$ values. The shaded green region illustrates the 16th to 84th percentiles around the median for the \coco{} sample. We also show the best fit power laws (Eq. \ref{eqn:vmax-msub}) to the \coco{} and \MII{} data.}

\caption{A comparison of the $V_{max}-R_{max}$ relation for haloes versus that for subhaloes. The solid yellow and orange lines show the relation for main haloes in the \coco{} and \scolor{} samples. The remaining lines show the $V_{max}-R_{max}$ relation for subhaloes identified in all the \coco{} hosts (green solid line); in Milky-Way mass hosts ( blue dashed line;see \S\ref{subsec:vmax-rmax}); and in the \aquarius{} level 2 hosts (solid brown line). The green shaded region gives the 16th to 84th percentiles for the full sample of \coco{} subhaloes. }

111 — 1505.06582

\caption{Circuit-like description of the \textcolor{blue}{proposed} generators.}

112 — 1505.06726

\caption{(color online) Metastable states of the stationary BoS game. In the mean-filed limit $N = \infty$, a trajectory spirals towards a fixed point $\left(\frac{1}{2},\frac{1}{2}\right)$, the Nash equilibrium of the game. For the finite $N$, metastable states are specified by their quasi-stationary probability density functions (pdf's) (3d plots). %With the increase of $N$, the functions %tend to localize at the Nash equilibrium. %Although, for any $N$, the mean position $\left(\bar{x}(t), \bar{y}(t)\right)$ coincide with the Nash equilibrium, %the stochastic evolution is governed by the metastable limit cycles located on the crater ridge %on the pdf's tops. For $N=200$ the pdf combines the results of the direct diagonalization of the $39~ 601 \times 39~ 601$ matrix $\tilde{\mathbf{Q}}$ (left half of the pdf, this procedure was also used to obtain the function for $N=100$) and of the preconditioned stochastic sampling (right part of the pdf, this procedure was also used to obtain the function for $N=400$) \cite{generic}. The baseline fitness $w = 0.3$ (other parameters are given in the text).}\label{Fig:2} \end{figure*} %------------------------------------------------------------------ We are interested in the dynamics before the fixation, so we merge the four states into a single absorbing state by summing the corresponding incoming rates. The boundary states, $(i = \{0,N\},j\in \{1,\cdots,N-1\})$ and $(i\in \{1,\cdots,N-1\},j=\{0,N\})$, can also be merged into this absorbing super-state: Once the population gets to the boundary, it will only move towards one of the two nearest absorbing states. By labeling the absorbing super-state with index $k = 0$, we end up with a $(L+1) \times (L+1)$ matrix \begin{eqnarray} \tilde{\mathbf{S}}^{m} = \begin{bmatrix} 1 & \boldsymbol{\varrho_0}^{m} \\ \mathbf{0} & \tilde{\mathbf{Q}}^{m}, \end{bmatrix} \label{Eq:supermatrix} \end{eqnarray} where $L = (N-1)^2$, $\boldsymbol{\varrho}_0^{m}$ is a vector of the incoming transition probabilities of the absorbing super-state, $\mathbf{0}$ is a $L\times 1$ zero vector, and $\tilde{\mathbf{Q}}^{m}$ is a $L \times L$ reduced transition matrix. With Eq.~(\ref{Eq:supermatrix}), we arrive at the setup used by Darroch and Seneta to formulate the concept of quasi-stationary distributions \cite{darroch}. There is the normalized right eigenvector of the reduced transition matrix $\tilde{\mathbf{Q}}^{m}$ with the maximum eigenvalue $\lambda$ \cite{FPT}. By using the inverse bijection, we can transform this vector %$\tilde{\mathbf{d}}$ into a two-dimensional probability density function (pdf), i.e., a state, $\mathbf{d}$, with maximal mean absorption time. This state is the most resistant to the wash-out by the finite-size fluctuations and it remains near invariant, up to a uniform rescaling, under the action of the tensor ${\mathbf{S}}$. This is the metastable state of the evolutionary process. \textit{Stationary case.} As an example, we consider a game with payoffs $a_{11}$, $a_{22}$, $b_{12}$ and $b_{21}$ equal $1$, and payoffs $-1$ for the rest of strategies \cite{neumann}. Figure 2 presents the numerically obtained metastable states of the game. We use two methods, the direct diagonalization of the reduced transition matrix, which is stationary in this case, $\tilde{\mathbf{Q}}^{m} \equiv \tilde{\mathbf{Q}}$, and preconditioned stochastic sampling \cite{generic}. %The latter was performed by %launching trajectories from random initial points, %uniformly distributed on the $N-1 \times N-1$ %grid and then sampling the pdf with only those trajectories which remained unabsorbed after $10\cdot N^2$ rounds. For $N=200$ we find an agreement between the results of the two methods. The means of the metastable state, \begin{eqnarray} \bar{x} = \sum_{i,j=1}^{N-1} \frac{i}{N} \cdot d(i,j) ;~\bar{y} = \sum_{i,j=1}^{N-1} \frac{j}{N} \cdot d(i,j), \label{Eq:mean} \end{eqnarray} coincide with the Nash equilibrium \cite{nash} for any $N$. However, the actual dynamics is determined by the metastable limit cycle encircling the equilibrium (this could be seen by performing short-run stochastic simulations); see Fig. 2. Within the Langevin-oriented approach to the dynamics of finite populations \cite{traulsen1,traulsen4}, the appearance of the metastable limit cycle can be interpreted as a stochastic Hopf bifurcation \cite{arnold} (see also Ref.~\cite{lindner} for another interpretation). In the limit $N \rightarrow \infty$ the cycle collapses to the Nash equilibrium. Note, however, that the convergence to this limit is slow, as indicated by the width of the pdf for $N=400$. \textit{Case of modulated payoffs}. By adding time-modulations to the model, we find that the mean-field dynamics does not exhibit substantial changes. For the choice $\epsilon(t) = \tilde{a}_{11}(t) = \tilde{b}_{22}(t)= f \cos(\omega t)$ with $\omega = 2\pi/T$ (all other payoffs held stationary) we observed a period-one limit cycle localized near the Nash equilibrium of the stationary case, see Figs.~3(a,b). It collapses to a set of adiabatic Nash equlibria, $\left\{x_{NE}(\epsilon) = \frac{2-\epsilon}{4 - \epsilon}, y_{NE}(\epsilon) = \frac{2}{4 + \epsilon}\right\}$ in the limit $\omega \rightarrow 0$. The dynamics of a finite $N$ population is different. The stochastic evolution of a trajectory in the $(i,j)$-space, initiated away from the absorbing boundary, can be divided into two stages. At first the trajectory relaxes towards a metastable state. The timescale of this process is defined by the mixing time $t_{\mathrm{mix}}(N)$ \cite{traulsen5}, which in this case has to be calculated now for the \textit{quasi}-stationary state. Then the trajectory wiggles around the metastable state until the fluctuations drive it to the absorbing boundary. Following the random-walk approximation, the mean absorption time $t_{\mathrm{abs}}(N)$, called ``mean fixation time'' \cite{dawkins,smith} in the evolutionary context, seemingly should also scale as $N$. However, this estimate neglects the presence of the inner attractive manifold and the fact that the noise strength decreases upon approaching the absorbing boundary. In fact, the absorption time scales super-linearly with $N$ \cite{meantime}. The lifetime of the metastable state is restricted to the time interval $[t_{\mathrm{mix}}(N), t_{\mathrm{abs}}(N)]$, whose length scales as $t_{\mathrm{abs}}(N)[1 - t_{\mathrm{mix}}(N)/t_{\mathrm{abs}}(N)] \sim t_{\mathrm{abs}}(N)$. For $\omega = 0.1$ \cite{size_issue}, the stochastic simulations reveal a metastable state which is distinctively different from the limit cycle produced by the mean-field equations, see Fig.~3b. There is a conflict between the evolution of means, described by the adjusted replicator equations, and the results of the stochastic dynamics. The conflict can be resolved with the concept of the quasi-stationary distribution. Namely, the transition matrices, Eq.~(\ref{Eq:supermatrix}), are round-specific now and form a set $\{\tilde{\mathbf{S}}^{m}\}$, $m=1,\cdots, M$. %(recall that after $M=T/\triangle t$ rounds the periodically modulated payoffs return to their initial value). The propagator over the time interval $[0,t]$, $0 < t < T$, is the product $\tilde{\mathbf{U}}(t) = \prod_{m'=1}^{M_{t}}\tilde{\mathbf{S}}^{m'}$ with $M_{t}=t/\triangle t$. All the propagators, including the period-one propagator $\tilde{\mathbf{U}}(T)$, have the same structure as the super-matrix in Eq.~(\ref{Eq:supermatrix}). We define the metastable state $\mathbf{d}(T)$ as the the quasi-stationary distribution of $\tilde{\mathbf{U}}(T)$. It is also a \textit{Floquet} state \cite{floquetH} of the reduced propagator $\tilde{\mathbf{U}}^{r}(T)$, which can be obtained by replacing the transition matrices $\tilde{\mathbf{S}}^{m'}$ with the matrices $\tilde{\mathbf{Q}}^{m'}$ or by simply cutting out the first line and column from the matrix $\tilde{\mathbf{U}}(T)$. The Floquet state is a time-periodic state, $\mathbf{d}(t + T) = \mathbf{d}(t)$, which changes during one period of modulations, see Fig.~4. The metastable state $\mathbf{d}(t)$ at any instant of time $t$, $0 < t <T$, can be obtained by acting on the state $\mathbf{d}(0)$ with the reduced propagator $\tilde{\mathbf{U}}^{r}(t)$. %--------------------------------------------------------------- \begin{figure}[t] %\includegraphics[width=0.45\textwidth]{fig2} \includegraphics[width=0.49\textwidth]{fig3n} \caption{(color online) Evolutionary dynamics governed by the BoS game with modulated payoffs. (a) Period-one limit cycles of the mean-field dynamics for $\omega = 0.1$ (blue dash-dotted line) and $\omega = 0.01$ (red solid line) are localized near the Nash equilibrium of the stationary game, $\left(\frac{1}{2},\frac{1}{2}\right)$ (arrows indicate the direction of motion). In the limit $\omega \rightarrow 0$, the mean-field attractor shrinks to the set of adiabatic Nash equlibria (black dashed line). Mean position ({\color{blue}$\bullet$}) of a finite-$N$ metastable Floquet state, $\left(\bar{x}(t), \bar{y}(t)\right)$, Eq.~(\ref{Eq:mean}), moves along the limit cycle localized near the point $\left(\frac{1}{2},\frac{1}{2}\right)$ (the means are plotted at the instants $t_n = nT/5$, $n=0,..,4$); (b) A stochastic trajectory (grey line) reveals the existence of a period-two limit cycle [the period doubling can be resolved with stroboscopic points, plotted at the instants $2nT$ ($\bigtriangleup$) and $(2n+1)T$ ($\diamondsuit$)]. The trajectory is initiated at the point marked with the open blue square and ends up at the absorbing state (red cross at the upper left corner). The trajectory of the mean of the finite-$N$ metastable Floquet state ({\color{blue}$\bullet$}) is distinctively different from the stochastic trajectory [note the change of scale as compared to panel (a)]. The parameters are $f = 0.5$, $N=200$, and $M=T/\triangle t=10N$ (corresponds to the driving frequency $\omega=0.1$ in the mean-field limit) \cite{size_issue}. Other parameters as in Fig. 2.}\label{Fig:3} \end{figure} %--------------------------------------------------------------- The evolution of the means of the pdf $\mathbf{d}(t)$ (see Fig.~4a), $\left(\bar{x}(t), \bar{y}(t)\right)$, is close to the period-one limit cycle, see blue dots on Fig.~3a. However, the Floquet state consists of two peaks produced by the noised period-two limit cycle (compare also the positions of the stroboscopic points in Fig.~3b with the pdf for $t=0$ in Fig.~4a). The peak contributions balance each other thus reducing the dynamics of the means to the vicinity of the the point $\left(\frac{1}{2},\frac{1}{2}\right)$. %--------------------------------------------------------------- \begin{figure}[t] %\includegraphics[width=0.45\textwidth]{fig2} \includegraphics[width=0.45\textwidth]{fig4n} \caption{(color online) Evolution of the metastable Floquet state over one period of modulations. The pdfs obtained by the direct diagonalization of the reduced period-one propagator for $N=200$. The corresponding means $\left(\bar{x}(t), \bar{y}(t)\right)$ are shown on Fig. 3a ({\color{blue} $\bullet$}). Plots for $t = 0$ (above the diagonal) and $t = T$ (below the diagonal) present the results of the stochastic sampling. %(b) The lifetime $t_{\mathrm{life}}$ as a function of the modulation strength $f$, %for the population size $N = 50$ ($\bigcirc$), $100$ ($\Box$), and $200$ ($\triangle$). %Other parameters are as in Fig. 2. }\label{Fig:4} \end{figure} %--------------------------------------------------------------- The lifetime of the state $\mathbf{d}(t)$ can be estimated with the largest eigenvalue $\lambda_T$, $0<\lambda_T<1$, of the matrix $\tilde{\mathbf{U}}^{r}(T)$. To compare it with lifetimes of stationary metastable states, we introduce the mean single-round exponent, $\bar{\lambda}_T = \lambda_T^{1/M}$ and define the mean lifetime as $t_{\mathrm{life}} = 1/(1-\bar{\lambda}_T)$ \cite{generic}. %Figure 4b shows the dependence of $t_{\mathrm{life}}$ on the strength of modulations. Aside of the slow decay trend, we found the effect of modulations not being strong. This is in stark contrast to the structure of the metastable states. Namely, while in the stationary limit the pdf $\mathbf{d}$ is localized near the Nash equilibrium, at the maximal distance from the absorbing boundaries, the metastable Floquet state is localized near the absorbing boundary, see Fig. 4. We also detect the increase of the boundary localization with the increase of the population size beyond $N = 200$. This suggests that, in the limit $N \rightarrow \infty$, the dynamics of the system is governed by a period-two limit cycle localized near the absorbing boundary. The boundary localization of the metastable attractor can be interpreted as the presence of small fractions of mutants \cite{smith}, i.e. the players that are using strategies different from that used by the majority of populations. The evolutionary dynamics of the mutant fractions looks like a repeating sequence of population bottlenecks \cite{dawkins,bottleneck} yet this only weakly affects fraction lifetimes \cite{issue1} even in the case of finite $N$. \textit{Conclusions.} We presented a concept of metastable Floquet states in game-driven populations when mate selection preferences are periodically changing in time. Here we combined the Floquet formalism with the concept of quasi-stationary distributions to reveal the existence of complex, liquid-like nonequlibrium dynamics in the strategy space which cannot be resolved within the mean-field framework. %By combining the Floquet ideology with the concept of quasi-stationary %distributions, we reveal the existence of complex nonequlibrium dynamics which cannot be resolved within %the mean-field framework. Metastable Floquet states are not restricted to the field of ecology studies but can emerge in different periodically modulated systems with stochastic event-driven dynamics. They may, for example, underlay a gene expression in a single cell, which is modulated by a circadian rhythm \cite{cyrcadian} and can provide new interpretations of the Bose-Einstein condensation in ac-driven atomic ensembles \cite{ketz,frey2}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{thebibliography}{1000} \bibitem{sexual} M. Andersson, \textit{Sexual Selection}(Princeton Univ. Press. Princeton, 1994). \bibitem{dawkins} R. Dawkins, \textit{The Selfish Gene} (Oxford University Press, Oxford, 1976). \bibitem{generic} See Supplemental Material for more information. \bibitem{lizard1} D. Crews, Science \textbf{189}, 1059 (1975). \bibitem{lizard2} M. M. Holmes, C. L. Bartrem, and J. Wade, Physiol. and Behav. \textbf{91}, 601 (2007). \bibitem{srev} V. D. Jennions and M. Petrie, Biol. Rev. Cambridge Philos Soc., \textbf{72} (2006). \bibitem{darroch} J. N. Darroch and E. Seneta, J. Appl. Prob. \textbf{2}, 88 (1965). \bibitem{floquet} G. Floquet, Annales de l'\'{E}cole Normale Sup\'{e}rieure \textbf{12}, 47 (1883). \bibitem{floquetH} M. Grifoni and P. H\"{a}nggi, Phys. Rep. \textbf{304}, 229 (1998). \bibitem{meta1} G. Biroli and J. Kurchan, Phys. Rev. E \textbf{64}, 016101 (2001). \bibitem{meta2} S. Rulands, T. Reichenbach, and E. Frey, J. Stat. Mech. L01003 (2011). \bibitem{assaf} M. Assaf and M. Mobilia, Phys. Rev. Lett. \textbf{109}, 188701 (2012). \bibitem{hof2} J. Hofbauer and K. H. Schlag, J. of Evol. Economics \textbf{10}, 523 (2000). \bibitem{schlag} K. H. Schlag, J. of Econom. Theory \textbf{78}, 130. \bibitem{traulsen1} A. Traulsen, J. C. Claussen, and C. Hauert, Phys. Rev. Lett. \textbf{95}, 238701 (2005). \bibitem{traulsen2} Ch. S. Gokhale and A. Traulsen, Dyn. Games and Appl. \textbf{4}, 468 (2014). \bibitem{traulsen3} A. Traulsen, J. C. Claussen, and C. Hauert, Phys. Rev. E. \textbf{74}, 011901 (2006). \bibitem{frey} A. Dobrinevski and E. Frey, Phys. Rev. E \textbf{85}, 051903 (2012). %\bibitem{pom} A. Pomiankowski, J. Theor. Biol. 128, 195 (1987). \bibitem{moran} P. A. P. Moran, \textit{The Statistical Processes of Evolutionary Theory} (Clarendon, Oxford, 1962). \bibitem{moranNature} M. A. Nowak, A. Sasaki, C. Taylor, and D. Fudenberg, Nature \textbf{428}, 646 (2004). %\bibitem{imitation} These two consecutive steps, death and birth, can %be reinterpreted as a single step of \textit{imitation}, i.e. adoption of the strategy of the %first player by the second one \cite{hof2,schlag}. \bibitem{smith} J. M. Smith, \textit{Evolution and the Theory of Games}(Cambridge University Press, Cambridge, 1982). %\bibitem{transition} Here we follow the convention that the stochastic matrix acts on %the probability column vector to the right. \bibitem{gant} E. Seneta, \textit{Non-negative Matrices and Markov Chains} (Springer, NY, 2006). \bibitem{dykman} M. Khasin and M. I. Dykman, Phys. Rev. Lett. \textbf{103}, 068101 (2009). \bibitem{FPT} By virtue of the Perron-Frobenius theorem, $\lambda$ and $\tilde{\mathbf{d}}$ are both real and non-negative \cite{gant}. \bibitem{neumann} This choice corresponds to the Matching Pennies game, see J. von Neumann and O. Morgenstern, \textit{Theory of Games and Economic Behaviour} (Princeton University Press, Princeton, 1944). \bibitem{nash} J. Nash, PNAS \textbf{36}, 48 (1950). \bibitem{traulsen4} A. Trauslen, J. C. Claussen, and C. Hauert, Phys. Rev. E \textbf{85}, 041901 (2012). \bibitem{arnold} L. Arnold, \textit{Random Dynamical Systems} (Springer, NY, 2003). \bibitem{lindner} P. J. Thomas and B. Lindner, Phys. Rev. Lett. \textbf{113}, 254101 (2014). %\bibitem{redner} S. Redner, \textit{A Guide to First Passage Processes} %(Cambridge University Press, Cambridge, 2001). \bibitem{traulsen5} A. J. Black, A. Traulsen, and T. Galla, Phys. Rev. Lett. \textbf{109}, 028101 (2012). \bibitem{meantime} The average absorption time for a specific initial state, $t_{\mathrm{abs}}(i,j)$, is proportional the corresponding entry in the \textit{left} maximal-eigenvalue eigenvector of the reduced matrix $\tilde{\mathbf{Q}}$. The proportionality coefficient can be found from the dual orthonormality condition. \bibitem{size_issue} We find a sharp contrast between the mean-filed dynamics and the stochastic evolution for this particular value of $\omega$. The optimal value for the frequency (period) of modulations could be different for other driving scheme and/or other choice of the game payoffs. \bibitem{bottleneck} T. Maruyama and P. A. Fuerst, Genetics \textbf{111}, 691 (1985). \bibitem{issue1} The relations between the exponent $\lambda_T$, mean absorption (fixation) time \cite{meantime}, and dynamical properties of Floquet states is an interesting issue. It can be explored, for example, with a discrete-time generalization of the ``optimal path to exctintion'' approach \cite{dykman2,escudero,meerson}. \bibitem{dykman2} M. I. Dykman, E. Mori, J. Ross, and P. M. Hunt, J. Chem. Phys. \textbf{100}, 5735 (1994). \bibitem{escudero} C. Escudero and J. A. Rodriguez, Phys. Rev. E \textbf{77}, 011130 (2008). \bibitem{meerson} M. Assaf, A. Kamenev, and B. Meerson, Phys. Rev. E \textbf{78}, 041123 (2008). %\bibitem{floquet_ecol1} C. A. Klausmeier, Theor. Ecol. \textbf{1}, 153 (2008); %C. T. Kremer and C. A. Klausmeier, J. Theor. Biol. \textbf{339}, 14 (2013). \bibitem{cyrcadian} S. S. Golden, V. M. Cassone, and A. Li Wang, Nat. Struct. Mol. Biol. \textbf{14}, 362 (2007); A. Sancar, Nat. Struct. Mol. Biol. \textbf{15}, 23 (2008). \bibitem{ketz} D. Vorberg, W. Wustmann, R. Ketzmerick, and A. Eckardt, Phys. Rev. Lett. \textbf{111}, 240405 (2013). \bibitem{frey2} J. Knebel, M. F. Weber, T. Kr\"{u}ger, and E. Frey, Nature Comm. \textbf{6}, 6977 (2015). \end{thebibliography} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{center} \textbf{Supplemental Material} \end{center} \section{Seasonal variations in mate preferences} Stable time variations were found in the female flycatcher preferences or male forehead patch size that resulted in late-breeding females preferring males with larger patches \cite{fly}. It was explained by the fact that in the beginning of the breeding season, large-patched males allocate more resources to courting than to parental care but change their habits to the opposite late in the season. Seasonal variations were also found in fiddle crabs (female preference to male claw size) \cite{crab}, two-spotted goby (female preference to overall male size) \cite{goby}, and sailfin mollies (male preferences for two different kind of females) \cite{molly}. %There is overall no %agreement reached between ecologists on the role the seasonal plasticity %in the mate preference (of different sexually selected traits) plays in %the determination of the evolution direction of a species, see Ref. \cite{srev}. \section{Moran process} Players $A$ (males) and $B$ (females) form two populations, each one of a fixed size $N$ and with two available strategies, $s = \{1,2\}$. Payoffs are specified by four functions, $\{a_{ss'}(t)\}$ and $\{b_{s's}(t)\}$, $s,s' = \{1,2\}$. The average payoff of the players using strategy $s$ is \begin{eqnarray} ~\pi^{A}_s(j,t) = a_{s1}(t)\frac{j}{N}+a_{s2}(t)\frac{(N-j)}{N}, \\ \pi^{B}_s(i,t) = b_{s1}(t)\frac{i}{N}+b_{s2}(t)\frac{(N-i)}{N}.~ \label{Eq:quasi_sym1} \end{eqnarray} Payoffs determine the probabilities for a player to be chosen for reproduction, e.g. for the male population, \begin{eqnarray} P^{A}_s(i,j,t) = \frac{1}{N} \cdot \frac{1 - w + w\pi^{A}_s(j,t)}{1 -w + w \bar{\pi}^{A}(i,j,t)}, \label{Eq:quasi_sym2} \end{eqnarray} where $\bar{\pi}^{A}(i,j,t) = [i\pi^{A}_{1}(j,t)+(N-i)\pi^{A}_{2}(j,t)]/N$ is the average payoff of the males. The baseline fitness $w \in [0,1]$ is a tunable baseline fitness parameter determining how the player's chance to be chosen for reproduction is related to player's performance \cite{moranNature,traulsen1}. When $w = 0$, the probability to be chosen for reproduction does not depend on player's performance and is uniform across the population. After the choice has been made, another member of the population is chosen completely randomly and replaced with an offspring of the player chosen for reproduction, i.e. with a player using the same strategy as its parent \cite{imitation}. This update mechanism is acting simultaneously in both populations, $A$ and $B$, such that a mating pair produces two offspring, a male and a female, on every round. Therefore, the size of the populations $N$ remains constant. %preserved and $N$ is a game parameter. A single round can be considered as a one-step Markov process, with transition rates, e.g. for population $A$, from a state $i$ to states $i+1$ and $i-1$, are given by \cite{traulsen1, traulsen4} \begin{eqnarray} \nonumber T_{A}^{+}(i,j,t) =\frac{1-w + w\pi^{A}_1(t)}{1 -w + w \bar{\pi}^{A}}\frac{i}{N}\frac{N-i}{N}, \\ T_{A}^{-}(i,j,t) =\frac{1-w + w\pi^{A}_2(t)}{1 -w + w \bar{\pi}^{A}}\frac{N-i}{N}\frac{i}{N}. \label{Eq:rates} \end{eqnarray} \section{Transition tensor} Here we describe the transition fourth-order tensor $S^{m}(i,j,i',j')$ in terms of the rates [$T_{A}^{+,-}(i,j,t)$ and $T_{B}^{+,-}(i,j,t)$] for populations $A$ and $B$ given by Eq.~(4) in the main text. The stochastic Moran process can be expressed as a Markov chain \cite{traulsen4} \begin{widetext} \begin{eqnarray} p^{m+1}(i,j) &=& \left[1-T_{A}^{+}(i,j,m\triangle t)-T_A^{-}(i,j,m\triangle t)\right]\left[1-T_B^{+}(i,j,m\triangle t)-T_B^{-}(i,j,m\triangle t)\right]p^{m}(i,j) \nonumber \\ && + T_B^{-}(i,j+1,m\triangle t)\left[1-T_A^{-}(i,j+1,m\triangle t)-T_A^{+}(i,j+1,m\triangle t)\right]p^{m}(i,j+1) \nonumber \\ && + T_B^{+}(i,j-1,m\triangle t)\left[1-T_A^{-}(i,j-1,m\triangle t)-T_A^{+}(i,j-1,m\triangle t)\right]p^{m}(i,j-1) \nonumber \\ && + T_A^{-}(i+1,j,m\triangle t)\left[1-T_B^{-}(i+1,j,m\triangle t)-T_B^{+}(i+1,j,m\triangle t)\right]p^{m}(i+1,j) \nonumber \\ && + T_A^{+}(i-1,j,m\triangle t)\left[1-T_B^{-}(i-1,j,m\triangle t)-T_B^{+}(i-1,j,m\triangle t)\right]p^{m}(i-1,j) \nonumber \\ && + T_A^{-}(i+1,j+1,m\triangle t)T_B^{-}(i+1,j+1,m\triangle t) p^{m}(i+1,j+1) \nonumber \\ && + T_A^{+}(i-1,j+1,m\triangle t)T_B^{-}(i-1,j+1,m\triangle t) p^{m}(i-1,j+1) \nonumber \\ && + T_A^{-}(i+1,j-1,m\triangle t)T_B^{+}(i+1,j-1,m\triangle t) p^{m}(i+1,j-1) \nonumber \\ && + T_A^{+}(i-1,j-1,m\triangle t)T_B^{+}(i-1,j-1,m\triangle t) p^{m}(i-1,j-1). \end{eqnarray} \end{widetext} The above equation can be recast into \begin{eqnarray} p^{m+1}(i,j) &=& \sum_{i',j'} S^{m}(i,j,i',j') p^{m}(i',j'), \end{eqnarray} where the fourth-order tensor $S^{m}(i,j,i',j')$ is given by, \begin{widetext} \begin{eqnarray} S^{m}(i,j,i',j') &=& \left[1-T_A^{+}(i',j',m\triangle t)-T_A^{-}(i',j',m\triangle t)\right]\left[1-T_B^{+}(i',j',m\triangle t)-T_B^{-}(i',j',m\triangle t)\right]\delta_{i',i}\,\delta_{j',j} \nonumber \\ && + T_B^{-}(i',j',m\triangle t)\left[1-T_A^{-}(i',j',m\triangle t)-T_A^{+}(i',j',m\triangle t)\right]\delta_{i',i}\,\delta_{j',j+1} \nonumber \\ && + T_B^{+}(i',j',m\triangle t)\left[1-T_A^{-}(i',j',m\triangle t)-T_A^{+}(i',j',m\triangle t)\right]\delta_{i',i}\,\delta_{j',j-1} \nonumber \\ && + T_A^{-}(i',j',m\triangle t)\left[1-T_B^{-}(i',j',m\triangle t)-T_B^{+}(i',j',m\triangle t)\right]\delta_{i',i+1}\,\delta_{j',j} \nonumber \\ && + T_A^{+}(i',j',m\triangle t)\left[1-T_B^{-}(i',j',m\triangle t)-T_B^{+}(i',j',m\triangle t)\right]\delta_{i',i-1}\,\delta_{j',j} \nonumber \\ && + T_A^{-}(i',j',m\triangle t)T_B^{-}(i',j',m\triangle t)\delta_{i',i+1}\,\delta_{j',j+1} \nonumber \\ && + T_A^{+}(i',j',m\triangle t)T_B^{-}(i',j',m\triangle t)\delta_{i',i-1}\,\delta_{j',j+1} \nonumber \\ && + T_A^{-}(i',j',m\triangle t)T_B^{+}(i',j',m\triangle t)\delta_{i',i+1}\,\delta_{j',j-1} \nonumber \\ && + T_A^{+}(i',j',m\triangle t)T_B^{+}(i',j',m\triangle t)\delta_{i',i-1}\,\delta_{j',j-1}. \end{eqnarray} \end{widetext} Above $i= 0, \cdots, N$, $j= 0, \cdots, N$, $i'= 0, \cdots, N$, and $j'= 0, \cdots, N$. Using the bijection $k=(N-1)j+i$ and $l=(N-1)j'+i'$, we obtain the required matrix form, see Eq.~(7) in the main text. \section{Adjusted replicator equations} In the continuous limit $N \rightarrow \infty$, the dynamics of the variables $x = i/N$ and $y = j/N$ is defined by the adjusted replicator equations \cite{smith,hof2}, \begin{align} &\dot{x} = [1-x][\Delta^A(t) - \Sigma^A(t)y]\frac{1}{\Gamma + \bar{\pi}^A(x,y,t)}, \\ &\dot{y} = [1-y][\Delta^B(t) - \Sigma^B(t)x]\frac{1}{\Gamma + \bar{\pi}^B(x,y,t)}, \label{Eq:replicator} \end{align} where $\Delta^{C} = c_{12} - c_{22}$, $\Sigma^{C}=c_{11}+c_{22}-c_{12}-c_{21}$, $\Gamma = \frac{1-w}{w}$, and $C=\{A,B\}$. $\bar{\pi}^A(x,y,t)$ [$ \bar{\pi}^B(x,y,t)$] is the averaged (over the population) payoff of the males [females]. \section{The lifetime of a metastable state} The lifetime of the state $\mathbf{d}(t)$ can be estimated with the largest eigenvalue $\lambda_T$, $0<\lambda_T<1$, of the matrix $\tilde{\mathbf{U}}^{r}(T)$. To compare it with lifetimes of stationary metastable states, we introduce the mean single-round exponent, $\bar{\lambda}_T = \lambda_T^{1/M}$ and define the mean lifetime as $t_{\mathrm{life}} = 1/(1-\bar{\lambda}_T)$. Figure 1 shows the dependence of $t_{\mathrm{life}}$ on the strength of modulations. %--------------------------------------------------------------- \begin{figure}[t] %\includegraphics[width=0.45\textwidth]{fig2} \includegraphics[width=0.45\textwidth]{figSupp} \caption{(color online) The lifetime $t_{\mathrm{life}}$ as a function of the modulation strength $f$, for the population size $N = 50$ ($\bigcirc$), $100$ ($\Box$), and $200$ ($\triangle$). Other parameters are as in Fig. 2 in the main text. }\label{Fig:1} \end{figure} %--------------------------------------------------------------- \section{Simulations} The preconditioned stochastic sampling was performed by launching trajectories from random initial points, uniformly distributed on the $N-1 \times N-1$ grid and then sampling the pdf with only those trajectories which remained unabsorbed after $10\cdot N^2$ rounds. For $N=200$ the diagonalization of the $39~ 601 \times 39~ 601$ matrix $\tilde{\mathbf{Q}}_T$ was performed on the cluster of the MPIPKS (Dresden) and Leibniz-Rechenzentrum (M\"{u}nchen). The stochastic sampling was performed on a GPU cluster consisting of twelve TESLA K20XM cards. That allowed us to obtain $5 \cdot 10^8$ realizations for each set of parameters. %For $N=200$ we find a perfect agreement between the results of the two approaches. \begin{thebibliography}{1000} \bibitem{fly} A. Qvarnstr\"{o}m, T. P\"{a}rt, and B. C. Sheldon, Nature \textbf{405}, 344 (2000). \bibitem{crab} R. N. C. Milner, \textit{et al.}, Behav. Ecology \textbf{21}, 311(2010). \bibitem{goby} A. A. Borg, E. Forsgren, and T. Amudsen, Anim. Behav. \textbf{72}, 763 (2006). \bibitem{molly} K. U. Heubel and J. Schlupp, Behav. Ecol. 19, 1080 (2008). %\bibitem{srev} V. D. Jennions and M. Petrie, Biol. Rev. Cambridge Philos Soc., \textbf{72} %(2006). %\bibitem{moran} P. A. P. Moran, \textit{The Statistical Processes of Evolutionary Theory} (Clarendon, Oxford, 1962). \bibitem{moranNature} M. A. Nowak, A. Sasaki, C. Taylor, and D. Fudenberg, Nature \textbf{428}, 646 (2004). \bibitem{traulsen1} A. Traulsen, J. C. Claussen, and C. Hauert, Phys. Rev. Lett. \textbf{95}, 238701 (2005). \bibitem{imitation} These two consecutive steps, death and birth, can be reinterpreted as a single step of \textit{imitation}, i.e. adoption of the strategy of the first player by the second one \cite{hof2,schlag}. \bibitem{hof2} J. Hofbauer and K. H. Schlag, J. of Evol. Economics \textbf{10}, 523 (2000). \bibitem{schlag} K. H. Schlag, J. of Econom. Theory \textbf{78}, 130. \bibitem{traulsen4} A. Trauslen, J. C. Claussen, and C. Hauert, Phys. Rev. E \textbf{85}, 041901 (2012). \bibitem{smith} J. M. Smith, \textit{Evolution and the Theory of Games}(Cambridge University Press, Cambridge, 1982). %\bibitem{Trauslen} A. Trauslen, J. C. Claussen, and C. Hauert, Phys. Rev. E \textbf{85}, 041901 (2012). \end{thebibliography} \end{document} }

\caption{(color online) Evolutionary dynamics governed by the BoS game with modulated payoffs. (a) Period-one limit cycles of the mean-field dynamics for $\omega = 0.1$ (blue dash-dotted line) and $\omega = 0.01$ (red solid line) are localized near the Nash equilibrium of the stationary game, $\left(\frac{1}{2},\frac{1}{2}\right)$ (arrows indicate the direction of motion). In the limit $\omega \rightarrow 0$, the mean-field attractor shrinks to the set of adiabatic Nash equlibria (black dashed line). Mean position ({\color{blue}$\bullet$}) of a finite-$N$ metastable Floquet state, $\left(\bar{x}(t), \bar{y}(t)\right)$, Eq.~(\ref{Eq:mean}), moves along the limit cycle localized near the point $\left(\frac{1}{2},\frac{1}{2}\right)$ (the means are plotted at the instants $t_n = nT/5$, $n=0,..,4$); (b) A stochastic trajectory (grey line) reveals the existence of a period-two limit cycle [the period doubling can be resolved with stroboscopic points, plotted at the instants $2nT$ ($\bigtriangleup$) and $(2n+1)T$ ($\diamondsuit$)]. The trajectory is initiated at the point marked with the open blue square and ends up at the absorbing state (red cross at the upper left corner). The trajectory of the mean of the finite-$N$ metastable Floquet state ({\color{blue}$\bullet$}) is distinctively different from the stochastic trajectory [note the change of scale as compared to panel (a)]. The parameters are $f = 0.5$, $N=200$, and $M=T/\triangle t=10N$ (corresponds to the driving frequency $\omega=0.1$ in the mean-field limit) \cite{size_issue}. Other parameters as in Fig. 2.}\label{Fig:3} \end{figure} %--------------------------------------------------------------- The evolution of the means of the pdf $\mathbf{d}(t)$ (see Fig.~4a), $\left(\bar{x}(t), \bar{y}(t)\right)$, is close to the period-one limit cycle, see blue dots on Fig.~3a. However, the Floquet state consists of two peaks produced by the noised period-two limit cycle (compare also the positions of the stroboscopic points in Fig.~3b with the pdf for $t=0$ in Fig.~4a). The peak contributions balance each other thus reducing the dynamics of the means to the vicinity of the the point $\left(\frac{1}{2},\frac{1}{2}\right)$. %--------------------------------------------------------------- \begin{figure}[t] %\includegraphics[width=0.45\textwidth]{fig2} \includegraphics[width=0.45\textwidth]{fig4n} \caption{(color online) Evolution of the metastable Floquet state over one period of modulations. The pdfs obtained by the direct diagonalization of the reduced period-one propagator for $N=200$. The corresponding means $\left(\bar{x}(t), \bar{y}(t)\right)$ are shown on Fig. 3a ({\color{blue} $\bullet$}). Plots for $t = 0$ (above the diagonal) and $t = T$ (below the diagonal) present the results of the stochastic sampling. %(b) The lifetime $t_{\mathrm{life}}$ as a function of the modulation strength $f$, %for the population size $N = 50$ ($\bigcirc$), $100$ ($\Box$), and $200$ ($\triangle$). %Other parameters are as in Fig. 2. }

113 — 1505.06874

\caption{The movement of the probability distribution $\rho(x,t)$ of the undamped oscillator in the first half period. The calculation is evaluated by the parameter set: $m = 1$; $\hslash = 1$; $\omega = 1$. The peak of the initial distribution is at $x = y_0 = -1$. The curves pertain to: {\color{blue}$t = 0$} (blue); {\color{purple}$t = T/8$} (purple); {\color{brown}$t = T/4$} (brown); {\color{green}$t = 3T/8$} (light green) and {\color{RoyalBlue}$t = T/2$} (light blue).}

\caption{The $\gamma$-governed shape change during the movement of the probability distribution $\rho(x,t)$ of the undamped oscillator in the first half period. The calculation is evaluated by the parameter set: $m = 1$; $\hslash = 1$; $\omega = 1$ and $\gamma = 0.8$. The peak of the initial distribution is at $x = y_0 = -1$. The curves pertain to: {\color{blue}$t = 0$} (blue); {\color{purple}$t = T/8$} (purple); {\color{brown}$t = T/4$} (brown); {\color{green}$t = 3T/8$} (light green) and {\color{RoyalBlue}$t = T/2$} (light blue).}

\caption{The change of the shape of distribution $\rho(x,t)$ in the first half period. The calculation is evaluated by the parameter set: $m = 1$; $\hslash = 1$; $\omega_r = 1$, $\hat\omega = 0.05$ and $\gamma = 0.8$. The initial position is at $x = y_0 = -1$. The {\color{blue}blue} curve on the left hand side pertains to the initial condition. The equidistant time steps between the {\color{purple}purple} -- {\color{brown}brown} -- {\color{green}light green} -- {\color{RoyalBlue}light blue} curves are ${\pi}/({4\sqrt{\omega_r^2 - \omega_c^2}})$.}

114 — 1505.06902

\caption{(a) Period and period spacings for Model\,4 (filled circles) and Model 1 (empty squares) from Table\,\ref{t-chisq}. The observed pattern is shown in grey. (b) The frequency difference between model frequencies and those detected in the observations $\delta f_i$ from Eq.\,(\ref{e-delta-f}). Compared to Fig.\,\ref{f-obs-dP}b, the ordinate is enlarged 40 times; the grey band around zero is the 1$\sigma$ frequency uncertainty range $\sigma_i$. A similar plot for Model\,4, Model\,8 and Model\,11 is shown in Figs.\,\ref{f-dP-4-vs-11} and \ref{f-dP-8-vs-11} in the Appendix. \label{f-best-chisq}}

\caption{Correlation diagram of the parameters of the Fine Grid around Model\,4. Results are based on the Mixing Grid plus the Fine Grid. The colour coding is based on$\log\chi^2_{\rm red}$ {\color{blue}}. To see the improvement we achieved in confining the grid parameter space, compare with Fig.\,11 in P14. In each panel, the model parameters that are not on the axis are fixed to those of Model\,4. Here,$X_{\rm c}$, $f_{\rm ov}$ and $\log D_{\rm mix}$ are well constrained.}

\caption{Brunt-V{\"a}is{\"a}l{\"a} frequency for Model\,4 (black solid), Model\,8 (red dashed), Model\,10 (grey solid) and Model\,11 (blue dotted) from Table\,\ref{t-chisq} and Table\,\ref{t-Model-4}. The inset is a zoom-in around the peak in the profile induced by the gradient of the mean molecular weight outside the convective core. These four models have different masses of the convective core M$_{\rm cc}$.}

\caption{Non-adiabatic instability study of Model\,4 (black circles), Model\,8 (red squares), and Model\,11 (blue stars) from Table\,\ref{t-chisq}. The imaginary part of the eigenfrequency $\omega_{\rm Im}$ is plotted against the mode period (vertical lines). Filled (empty) symbols designate excited (damped) modes.}

\caption{Top panel: Period spacing for Model\,4 (filled circles) and Model\,11 (empty squares). See Table\,\ref{t-chisq} for their parameters. Bottom panel: Absolute frequency deviations of the two models with respect to observations. See also Fig.\,\ref{f-best-chisq} for a comparison.}

\caption{Similar to Fig.\,\ref{f-best-chisq} but for Model\,8 (filled circles) and Model\,11 (empty squares).}

115 — 1505.07111

\caption{ \label{fig:fiducial1} Results of the simulation with the fiducial parameters. From top to bottom, the panels show the surface density $\Sigma$, velocity dispersion $\sigma$, inward radial velocity $-v_r$, instantaneous inflow rate $\dot{M}_{\mathrm{in}} = -2\pi r\Sigma v_r$, Toomre $Q$ parameter, and instability growth time normalized to the orbital time $t_{\mathrm{growth}}/t_{\mathrm{orb}}$. \red{In the top four panels, we show results at $t=0$ (\textit{thin black solid line}), $t=10$ Myr (\textit{thick blue solid line}), $t=25$ Myr (\textit{thick green dashed line}), and $t=50$ Myr (\textit{thin red dashed line}), as indicated in the legend.} For the instability growth time \red{in the bottom panel}, we show the value for the fastest-growing mode, with a solid line indicating that the fastest mode is acoustic and a dashed line indicating that it is gravitational. \red{Colors and line thicknesses are as in the panels above.} }

\caption{ \label{fig:alphastudy1} Results of the simulations with all parameters set to their fiducial values (\autoref{tab:fiducial}) except $\alpha_0$, which varies as indicated in the legend. The top row shows the surface density $\Sigma$ and the bottom shows $Q$; the left column shows results at $t=15$ Myr, and the right at $t=45$ Myr. \red{The fiducial case is shown as the solid thick black line, in this and all subsequent figures in this section.} }

116 — 1505.07192

\caption{ Parameter Selection and Model Robustness. We test different parameters on three benchmarks in terms of F-measure (higher is better) and MAE (lower is better). The best parameters are written in \textbf{bold}, which are our model's default settings. \textbf{\textcolor[rgb]{1.00,0.00,0.00}{Red}} and \textcolor[rgb]{0.00,0.00,1.00}{\textbf{blue}} numbers in bold represent the best and better performance in each evaluation category. %The best three results are highlighted %in \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green}, respectively. }

117 — 1505.07743

\caption{\small Statistics of the multi-frequency morphological differences inferred from Monte-Carlo (MC) simulations. All images have been restored from {\tt uvmem} models. \textcolor{blue}{\bf a)}: restored ATCA image at 34~GHz. \textcolor{blue}{\bf b)}: average of ATCA observations with 200 realizations of noise on the band~7 data. \textcolor{blue}{\bf c)}: average of the MC simulations for band~9 \textcolor{blue}{\bf d)-g)}: different realizations of noise. \textcolor{blue}{\bf h)}: rms scatter of the band~9 simulations. \textcolor{blue}{\bf i)}: rms scatter of the band~7 simulations. \label{fig:MCtests}}

\caption{\small Optical depth maps derived from the ATCA/ALMA multi-frequency data. The emissivity minimum at 11.5h indicates the presence of larger grains. The \textcolor{blue}{\bf upper panel} corresponds to the absolute flux density scale labelled {\em C13scale} in Sec.~\ref{sec:absoluteSED}, with a high flux density in band~7 (see Sec.~\ref{sec:absoluteSED}), while the \textcolor{blue}{\bf lower panel} corresponds to the flux scale in our preferred RADMC3D model, labelled {\em SEDscale} (see Sec.~\ref{sec:absoluteSED}). {\bf a}: optical depth map at the reference frequency of 345~GHz. {\bf b}: line of sight emissivity index map $\beta_s(\vec{x})$, with ATCA specific intensity contours in red. {\bf c}: root-mean-square uncertainties on the emissivity index map. {\bf d}: line of sight temperature, $T_s(\vec{x})$. \label{fig:Ttaubet} }

118 — 1505.07874

\caption{(Color online) X-ray diffraction data using CuK$_{\alpha}$ radiation for the single crystal used in this work. The crystallographic parameters are $a= 3.16(5) $ \AA \textcolor[rgb]{1,1,0.88}{9}and $c =3.626$ \AA.}

119 — 1505.07898

\caption{Time lines of the amplitudes $A_{max}$ ($\circ$) and $A_{min}$ (\textcolor{red}{$\vartriangle$}) with $T={\tilde t} \varepsilon^3$ for different values of $La$ and $\varepsilon$. The lines correspond to the exponential behavior $A=0.05 \, \exp[\Omega_m (T-T_0)]$, where $\Omega_m$ corresponds to the value given by either the 2D (solid line) or 1D (dashed line) model. $T_0 = 0$ except for the cases $(La,\varepsilon)=(10^4,0.5)$, $(10^2,1)$ and $(10^4,1)$.}

120 — 1505.08016

\caption{Analysis of electron acceleration in tightly-focused RPFLP as a function of the laser power. Red dots (\textcolor{red}{$\bullet$}) are 3DPIC data points from Fig. 5(a) of \cite{marceau2015_jpb}, while the solid black line (--) is a fit of the form $a + b\sqrt{P}$ (see also \cite{fortin10_jpb}). We identified in gray two different ultrashort electron pulse regimes: the subrelativistic slicing regime, where only a thin disk of electron is accelerated (see, e.g., \cite{marceau2013_prl}), and the relativistic bunching regime, where electrons are axially compressed into attosecond pulses (see, e.g., \cite{varin13_applsci}).\label{fig:scaling}}

121 — 1506.00278

\caption{All 12 types of Madlibs instructions and prompts. Right-most column shows the average number of words for each description (\textcolor{grey}{\#words for prompt} + \#words for answer).}

122 — 1506.00440

\caption{R/Q versus position for the first \colour{three modes M1, M2 and M3} from left to right, around the center of the beam pipe. Top row circular, bottom row elliptical. Color scale for R/Q is logarithmic. (color)}

\caption{\label{tab:tab}Eigenmode frequency and unloaded Q values$^{\dagger}$ \colour{of the first four modes.}}

123 — 1506.00594

\caption{\colorred \label{schematic}Partial phase diagram showing competing equilibrium R-R (black) and metastable AR-AR (red) phase coexistences \cite{Williamson2014}. Spinodals enclose the regions of local stability. Cartoons of the dominant inter-leaflet arrangement in each bilayer phase are shown. The grey dotted line illustrates R-AR coexistence, which is briefly discussed in Section~\ref{sec:def}. Other phase coexistences not considered here are omitted \cite{Williamson2014}. Parameters:\$\Delta_0 = 2\,a$, $\kappa = 3\,a^{-2}k_\textrm{B}T$, $V=0.6\,k_\textrm{B}T$, $J = 4\,a^{-2}k_\textrm{B}T$, $B=0.48\,a^{-2}k_\textrm{B}T$.}

\caption{\colorred \label{lattice}(a)~Mixed bilayer containing $S$ and $U$ ($L_o$ or gel and $L_d$-like) model species, illustrating the locally symmetric (R-R) and locally asymmetric (AR-AR) phase coexistences considered here. (b)~Microscopic lattice model for coupled leaflets, which can be coarse-grained (CG) (c) to give the mean-field free-energy density $f(\phi^\textrm{t},\phi^\textrm{b})$ as a function of locally-averaged leaflet compositions, and analysed for kinetics of domain formation with the inclusion of gradient costs for domain boundaries \cite{Williamson2014}. The lattice model can also be directly simulated.}

\caption{\colorred \label{kinetic}Red/blue colours and dashed/dotted lines:\mean-field theory parameter map showing relative growth rates of AR versus R instability modes for a bilayer comprising equimolar mixed leaflets, from linear stability analysis of initial demixing\cite{Williamson2014}, with $\Delta_0 = 2\,a$, $\kappa = 3\,a^{-2}k_\textrm{B}T$, $V=0.6\,k_\textrm{B}T$. The equilibrium state is R-R in all cases. Below the ``AR instability'' line AR-AR coexistence is possible but is metastable. Below the ``R instability'' line the homogeneous state is not unstable to the R mode, although R-R separation is still the equilibrium state. Overlaid dots from simulation (with $V=0.9\,k_\textrm{B}T$) show the average degree of registration $\lambda$ at the end of the simulation time, $\lambda = 0$ (black, AR-AR) to $\lambda = 1$ (white, R-R). We identify three kinetic classes discussed in the text (circled 1, 2, 3, bold line marks approximate boundaries). Illustrative simulation snapshot sequences for each class are shown ($\mathcal{L}\!=\!200$ in the snapshots). Simulations are visualised with OVITO \cite{OVITO}.}

124 — 1506.00993

\caption{Example of mass distributions in the circumprimary ring as a function of the asteroids' semi-major axis under the gravitational influence of a secondary M star at a$_{\scriptscriptstyle \text{b}}$ = 100 au and e$_{\scriptscriptstyle \text{b}}$ = 0.1 (\textcolor{gray}{$\blacksquare$}) and e$_{\scriptscriptstyle \text{b}}$ = 0.5 (\textcolor{black}{$\bullet$}). The vertical line refers to the gas giant's position. All particles are initially spaced by several Hill's radii to prevent immediate dynamical instability.}

\caption{Initial semi-major axis a$_{\scriptscriptstyle 0}$ and final periapsis distance q$_{\scriptscriptstyle {\text{HZc}}}$ of the incoming HZc for a single (\textcolor{black}{$\bullet$}) and binary (\textcolor{gray}{$\blacksquare$}) star system.}

125 — 1506.01144

\caption{BLEU-1,2,3,4, METEOR, CIDEr and $\mathcal{PPL}$ metrics compared with other state-of-the-art methods and our baseline on MS COCO dataset. $\ddagger$ indicates ground truth attributes labels are used, which (in \colorbox[rgb]{0.7,0.7,0.7}{gray}) will not participate in rankings.}

\caption{Results on the open-answer task for various question types on VQA (July Release Version). All results represent the percentage of answers in agreement with 10 human subjects, as calculated using the VQA evaluation tools \cite{antol2015vqa}. $\ddagger$ indicates that ground truth attributes labels are used. Results (in \textcolor{gray}{gray}) from \cite{antol2015vqa} are reported based on a different split, which contains 3 answers per question.}

\caption{Language generators for different types of tasks: (a) Image Captioning, (b) VQA-single word, (c) VQA-sentence. \textcolor[rgb]{1.00,0.00,0.00}{red} arrow indicates our attributes input $\Att(I)$ while \textcolor[rgb]{0.00,0.00,1.00}{blue} dash arrow shows the baseline method input $\CNN(I)$.}

126 — 1506.02446

\caption{{\it The CMA approach to approximating the effective generator $\mathcal{G}$ of the slow variables \blue{on the (possibly truncated) domain $S \in \Omega$,} without the need for stochastic simulations.}\label{CMA table}}

\caption{\it \blue{(a) A log plot of an approximation $\pi_\Omega$ of the invariant distribution on the slow variable $S = X_1 + 2X_2$ of system \eqref{eq:bis} with parameters \eqref{eq:bis:params}, demonstrating the bistable nature of the system. Approximation was computed by finding the null space of the full generator of the system on the truncated domain $\{0,1,\ldots,800\}\times\{0,1,\ldots,1200\}$. (b) Proportion of total propensity $P_{R_5,R_6}(X_1,X_2)$ attributed to the fast reactions $R_5$ and $R_6$, given by \eqref{eq:FPP}.\label{fig:bis}}}

127 — 1506.02993

\caption{Schematics of the photo-current model: the interface region $\Delta_0$ and the outer regions $\Delta_1$ and $\Delta_2$. \cblue{$\Sigma_s$ ($s$=1,2) represent planar interfaces $\Delta_s\cap\Delta_0$, located at $x_s$.} Wavepackets are generated in $\Delta_0$ and the charge separation is evaluated for each region at a typical coherence time $\tau_c$. }

128 — 1506.03212

\caption{\red{(a) Schematic diagram showing four equilibria for an example planar vector field. If we examine the three equilibria $\xi_1,\xi_2$ and $\xi_3$ there are heteroclinic connections from $\xi_2$ to $\xi_3$ and from $\xi_3$ to $\xi_1$. There is an excitable connection with threshold $\delta_{th}>0$ that corresponds to the radius of the black circle around $\xi_1$; for any $\delta>\delta_{th}$ (see for example the red circle) there is a connection shown in red from $\xi_1$ to $\xi_2$ with amplitude $\delta$. (b) and (c) show the excitable networks between the $\{\xi_i\}$ for amplitudes $\delta<\delta_{th}$ and $\delta>\delta_{th}$ respectively; note that the heteroclinic network between these equilibria is (b). The threshold corresponds to the distance of the stable manifold (shown by the dotted line in (a)), of the saddle equilibrium $\zeta$, from $\xi_1$.}}

\caption{\red{Schematic diagram illustrating the difference between a heteroclinic connection and an excitable connection with zero threshold. Four equilibria $\xi_i$ are such that there is an excitable (but not a heteroclinic) connection shown in red from any of the $\xi_i$, $i=1,2,3$ to $\xi_4$. Note that the alpha-limit set of the red trajectory contains the heteroclinic cycle between the $\xi_i$, $i=1,2,3$; it is a ``depth two'' connection \cite{ashwin_field_99}.}}

129 — 1506.03431

\caption{Transformations for \gls{ADVI}. The \textcolor{POSTcolor}{purple} line is the posterior. The \textcolor{Qcolor}{green} line is the approximation. \textbf{(a)} The latent variable space is $\bbR^+$. \textbf{(a$\to$b)} $T$ transforms the latent variable space to $\bbR$. \textbf{(b)} The variational approximation is a Gaussian. \textbf{(b$\to$c)} $S_{\mbmu,\mbomega}$ absorbs the parameters of the Gaussian. \textbf{(c)} We maximize the \gls{ELBO} in the standardized space, with a fixed standard Gaussian approximation.}

130 — 1506.03493

\caption{ \scriptsize The mode, arithmetic expectation, and geometric expectation of a Gamma-distributed random variable $\theta$. \emph{First}: The three quantities for different values of shape $a\geq 1$ (x axis) with rate $b=0.5$. All three grow linearly with $a$ and $\mathbbm{E}\left[\theta\right] \geq \mathbbm{G}\left[\theta\right] \geq \text{Mode}\,(\theta)$. \emph{Second}: Geometric and arithmetic expectations for different values of shape $a \in (0,1)$, where the mode is undefined, with rate $b=0.5$. $\mathbbm{G}\left[\theta\right]$ grows more slowly than $\mathbbm{E}\left[\theta\right]$. This property is most apparent when $a < 0.4$. \emph{Third}: pdf of a Gamma distribution with shape $a = 10$ and rate $b = 0.5$. The three quantities are shown as vertical lines. All three are close in the area of highest density, differing by about a half unit of inverse rate, i.e., $\frac{1}{2b} = 1$. \emph{Fourth}: pdf of a Gamma distribution with $a = 0.3$ and $b = 0.5$. The geometric and arithmetic expectations are shown as vertical lines (the mode is undefined). The two quantities differ greatly, with $\mathbbm{G}\left[\theta\right]$ much closer to zero and in an area of higher density. If these expectations were used as point estimates to predict the presence or absence of a rare event---e.g., $y = 0$ if $\hat{\theta} < 0.5$; otherwise $y = 1$---they would yield different predictions.} \vspace{-1em} \end{figure*} \section{Technical Discussion} \label{sec:discussion} Previous work on Bayesian Poisson matrix factorization (e.g.,~\cite{cemgil09bayesian, paisley14bayesian, gopalan15scalable}) presented update equations for the variational parameters in terms of auxiliary variables, known as \emph{latent sources}, and made no explicit reference to geometric expectations. In contrast, we write the update equations for Bayesian Poisson tensor factorization in the form of equations~(\ref{eqn:gamma_update}) and (\ref{eqn:delta_update}) in order to highlight their relationship to Lee and Seung's multiplicative updates for non-negative tensor factorization---a parallel also drawn by Cemgil in his paper introducing Bayesian PMF~\cite{cemgil09bayesian}---and to show that our update equations suggest a new way of making out-of-sample predictions when using BPTF. In this section, we provide a discussion of these connections and their implications. When performing NTF by minimizing the generalized KL divergence of reconstruction $\hat{\boldsymbol{Y}}$ from observed tensor $\boldsymbol{Y}$ (which is equivalent to MLE for PTF), the multiplicative update equation introduced by Lee and Seung for, e.g., $\theta^{(1)}_{ik}$ is \begin{equation} \label{eq:lee_seung} \theta^{(1)}_{ik} := \theta^{(1)}_{ik} \frac{\sum_{j,a,t} \theta^{(2)}_{jk} \theta^{(3)}_{ak} \theta^{(4)}_{tk}\, \frac{y_{ijat}}{\hat{y}_{ijat}} }{\sum_{j,a,t} \theta^{(2)}_{jk} \theta^{(3)}_{ak} \theta^{(4)}_{tk}}. \end{equation} These update equations sometimes converge to locally non-optimal values when the observed tensor is very sparse~\cite{gonzalez05accelerating,lin07convergence,chi12tensors}. This problem occurs when factors are set to \emph{inadmissible zeros}; the algorithm cannot recover from these values due to the multiplicative nature of the update equations. Several solutions have been proposed to correct this behavior when minimizing Euclidean distance. For example, Gillis and Glineur~\cite{gillis08nonnegative} add a small constant $\epsilon$ to each factor to prevent them from ever becoming exactly zero. For KL divergence, Chi and Kolda~\cite{chi12tensors} proposed an algorithm---Alternating Poisson Regression---that ``scooches'' factors away from zero more selectively (i.e., some factors are still permitted to be zero). In BPTF, point estimates of the latent factors are not estimated directly. Instead, variational parameters for each factor, e.g., $\gamma^{(1)}_{ik}$ and $\delta^{(1)}_{ik}$ for factor $\theta^{(1)}_{ik}$, are estimated. These parameters then define a Gamma distribution over the factor as in equation~(\ref{eq:q_for_single_theta}), thereby preserving uncertainty about its value. In practice, this approach solves the instability issues suffered by MLE methods, without any efficiency sacrifice. This assertion is supported empirically by the out-of-sample predictive performance results reported in section~\ref{sec:eval}, but can also be verified by comparing the form of the update in equation~(\ref{eq:lee_seung}) with those of the updates in equations~(\ref{eqn:gamma_update}) and (\ref{eqn:delta_update}). Specifically, if equations~(\ref{eqn:gamma_update}) and~(\ref{eqn:delta_update}) are substituted into the expression for the arithmetic expectation of a single latent factor, e.g., $\mathbb{E}\left[\theta^{(1)}_{ik} \right]= \frac{\gamma^{(1)}_{ik}}{\delta^{(1)}_{ik}}$, then the resultant update equation is very similar to the update in equation~(\ref{eq:lee_seung}): \begin{equation*} \mathbb{E}_{Q}\left[ \theta_{ik}^{(1)}\right] := \frac{\alpha + \sum_{j,a,t} \mathbb{G}_Q\left[ \theta^{(1)}_{ik} \theta^{(2)}_{jk} \theta^{(3)}_{ak} \theta^{(4)}_{tk} \right] \frac{y_{ijat}}{\hat{y}_{ijat}}}{\alpha\,\beta^{(1)} + \sum_{j,a,t} \mathbb{E}_Q \left[ \theta_{jk}^{(2)} \theta_{ak}^{(3)} \theta_{tk}^{(4)} \right]}, \end{equation*} where $\hat{y}_{ijat} \equiv \sum_{k=1}^K \mathbb{G}_Q \left[ \theta^{(1)}_{ik} \theta^{(2)}_{jk} \theta^{(3)}_{ak} \theta^{(4)}_{tk} \right]$. Pulling $\mathbb{G}_Q \left[ \theta^{(1)}_{ik}\right]$ outside the sum in the numerator and letting $\alpha \rightarrow 0$, yields \begin{equation*} \mathbb{E}_{Q}\left[ \theta_{ik}^{(1)}\right] := \mathbb{G}_Q \left[ \theta_{ik}^{(1)} \right] \frac{ \sum_{j,a,t} \mathbb{G}_Q\left[ \theta^{(2)}_{jk} \theta^{(3)}_{ak} \theta^{(4)}_{tk} \right] \frac{y_{ijat}}{\hat{y}_{ijat}} }{ \sum_{j,a,t} \mathbb{E}_Q \left[ \theta_{jk}^{(2)} \theta_{ak}^{(3)} \theta_{tk}^{(4)} \right]}, \end{equation*} which is exactly the form of equation~(\ref{eq:lee_seung}), except that the point estimates of the factors are replaced with two kinds of expectation. This equation makes it clear that the properties that differentiate variational inference for BPTF from the Lee and Seung updates for PTF are 1) the hyperparameters $\alpha$ and $\beta^{(1)}$ and 2) the use of arithmetic and geometric expectations of the factors instead of direct point estimates. Since the hyperparameters provide a form of implicit correction, BPTF should not suffer from inadmissible zeros, unlike non-Bayesian PTF. It is also interesting to explore the contribution of the geometric expectations. The fact that each $\hat{y}_{ijat}$ is defined in terms of a geometric expectation suggests that when constructing point estimates of the latent factors from the variational distribution (e.g., for use in prediction), the geometric expectation is more appropriate than the arithmetic expectation (which is commonly used in Bayesian Poisson matrix factorization) since the inference algorithm is implicitly optimizing the reconstruction as defined in terms of geometric expectations of the factors. To explore the practical differences between geometric and arithmetic expectations of the latent factors under the variational distribution, it is illustrative to consider the form of $\textrm{Gamma}\,(\theta; a, b)$. Most relevantly, the Gamma distribution is asymmetric, and its mean (i.e., its arithmetic expectation) is greater than its mode. When shape parameter $a \geq 1$, $\textrm{Mode}\,(\theta) = \frac{(a-1)}{b}$; when $a < 1$, the mode is undefined, but most of the distribution's probability mass is concentrated near zero---i.e., the pdf increases monotonically as $\theta \rightarrow 0$. This property is depicted in figure~2. In this scenario, the Gamma distribution's heavy tail pulls the arithmetic mean away from zero and into a region of lower probability. The geometric expectation is upper-bounded by the arithmetic expectation---i.e., $\mathbb{G}\left[\theta \right]= \frac{\exp{\left(\Psi(a) \right)}}{b} \leq \frac{a}{b} = \mathbb{E}\left[\theta \right]$. Unlike the mode, it is well-defined for $a \in (0,1)$ and grows quadratically over this interval, since $\exp{\left( \Psi(a) \right)} \approx \frac{a^2}{2}$ for $a \in (0,1)$; in contrast, the arithmetic expectation grows linearly over this interval. As a result, when $a < 1$, the geometric expectation yields point estimates that are much closer to zero than those obtained using the arithmetic expectation. When $a \geq 1$, $\exp{\left( \Psi(a) \right)} \approx a - 0.5$ and the geometric expectation is approximately equidistant between the arithmetic expectation and the mode---i.e., $\frac{a}{b} \geq \frac{a - 0.5}{b} \geq \frac{a-1}{b}$. These properties are depicted in figure~7; the key point to take away from this figure is that when $a < 1$, the geometric expectation has a much more probable value than the arithmetic expectation, while when $a \geq 1$, the geometric and arithmetic expectations are very close. This observation suggests that the geometric expectation should yield similar or better point estimates of the latent factors than those obtained using the arithmetic expectation. In table~2, we provide a comparison of the out-of-sample predictive performance for BPTF using arithmetic and geometric expectations. Indeed, these results show that the performance obtained using geometric expectations is either the same as or better than the performance obtained instead using arithmetic expectations. \begin{table}[t] \label{fig:ari-vs-geo} \scriptsize \caption{\scriptsize Predictive performance obtained using geometric and arithmetic expectations. (The experimental design was identical to that used to obtain the results in table~1.) Using geometric expectations resulted in the same or better performance than that obtained using arithmetic expectations.} \begin{center} \begin{tabular}{l l l l l l l} \toprule & & \multicolumn{2}{c}{BPTF-ARI} & \multicolumn{2}{c}{BPTF-GEO}\\ \cmidrule(r){3-4} \cmidrule(r){5-6} & Density & MAE & HAM-Z & MAE & HAM-Z\\ \midrule I-top-25 & 0.1217 &2.03 & 0.121 & {\bf 1.99} & {\bf 0.113}\\ %\midrule G-top-25 & 0.2638 &8.96 & 0.3 & {\bf 8.94} & {\bf 0.292}\\ %\midrule I-top-100 & 0.0264 &0.197 & 0.0236 & {\bf 0.178} & {\bf 0.0142}\\ %\midrule G-top-100 & 0.0588 &1 & 0.0857 & {\bf 0.95} & {\bf 0.0682}\\ \midrule I-top-25$^{c}$ & 0.0021 &0.0104 & 0.00163 & {0.0104} & {0.00161}\\ %\midrule G-top-25$^{c}$ & 0.0060 &0.0414 & 0.00606 & {\bf 0.0412} & {\bf 0.00601}\\ %\midrule I-top100$^{c}$ & 0.0004 &0.0011 & 5.03e-05 & {0.00109} & {4.97e-05}\\ %\midrule G-top100$^{c}$ & 0.0015 &0.00804 & 0.000959 & {0.00803} & {0.000957}\\ \bottomrule \end{tabular} \end{center} \end{table} \section{Summary} Over the past fifteen years, political scientists have engaged in an ongoing debate about using dyadic events to study inherently multilateral phenomena. This debate, as summarized by Stewart~\cite{stewart14latent}, began with Green et al.'s demonstration that many regression analyses based on dyadic events were biased due to implausible independence assumptions~\cite{green01dirty}. Researchers continue to expose such biases, e.g., ~\cite{erikson14dyadic}, and some have even advocated eschewing dyadic data on principle, calling instead for the development of multilateral event data sets~\cite{poast10misusing}. Taking the opposite viewpoint---i.e., that dyadic events can be used conduct meaningful analyses of multilateral phenomena---other researchers, beginning with Hoff~\cite{hoff04modeling}, have developed Bayesian latent factor regression models that explicitly model unobserved dependencies as occurring in some latent space, thereby controlling for their effects in analyses. This line of research has seen an increase in interest and activity over the past few years~\cite{hoff13equivariant,stewart14latent,hoff14multilinear}. In this paper, we too take this latter viewpoint, but instead of focusing on latent factor models for regression, we present a Bayesian latent factor model for predictive and exploratory data analysis---specifically, for identifying and characterizing the ``complex dependence structures in international relations''~\cite{king01proper} implicit in dyadic event data. Our exploratory analysis revealed interpretable multilateral structures that capture both persistent regional relations and temporally localized anomalies. As evidenced empirically by our predictive experiments and analytically by a comparison of our variational inference algorithm with traditional algorithms for performing non-negative tensor factorization, Bayesian Poisson tensor factorization overcomes the instability issues exhibited by standard non-negative tensor factorization methods when decomposing sparse, dispersed count data. We provided additional analysis and empirical results demonstrating that when constructing point estimates of the latent factors from the variational distribution, the geometric expectation is a more appropriate choice than the arithmetic expectation. We therefore recommend its use in any subsequent work involving variational inference for Bayesian Poisson matrix or tensor factorization. \section{Acknowledgments} Thank you to Mingyuan Zhou, Brendan O'Connor, Brandon Stewart, Roy Adams, David Belanger, Luke Vilnis, and Juston Moore for very helpful discussions. This work was partially undertaken while Aaron Schein was an intern at Microsoft Research New York City. This work was supported in part by the UMass Amherst CIIR and in part by NSF grants \#IIS-1320219,\#SBE-0965436, and\#IIS-1247664; ONR grant\#N00014-11-1-0651; and DARPA grant\#FA8750-14-2-0009. Any opinions, findings, and conclusions or recommendations expressed in this material are the authors' and do not necessarily reflect those of the sponsor.% % The following two commands are all you need in the % initial runs of your .tex file to % produce the bibliography for the citations in your paper. % \pagebreak \bibliographystyle{abbrv} \bibliography{ptf} % You must have a proper ".bib" file % and remember to run: % latex bibtex latex latex % to resolve all references % % ACM needs 'a single self-contained file'! % That's all folks! \end{document} }}

131 — 1506.04545

\caption{ (Color online) (a) Phase diagram of the effective Hamiltonian Eq.~(\ref{eq.1}) determined from CMC simulations. The shading (color) represents $T_{\text{c}}$. {\color[rgb]{0,0,0} Two quadrupole LRO phases, the planar antiferropseudospin (PAF) and planar ferropseudospin (PF) phases, exist in the vicinity of the SI phase~\cite{S.Onoda2011PRB}.} Classical SI is replaced by a U(1) QSL in quantum theory~\cite{LeePRB2012}. % The region enclosed by the dotted line represents an acceptable parameter region for the experimental data on Tb$_{2.005}$Ti$_{1.995}$O$_{7+y}$. % The cross mark indicates the typical values $(\delta, q) = (0,0.85)$. (b) Schematic view of the deformation of the $f$-electron charge density due to the PAF order on the pyrochlore lattice. }

\caption{ (Color online) (a) Temperature dependence of the observed specific heat $C_P(T, H)$ for $H\parallel[111]$. (b) Temperature-field map of $C_P(T, H)$ for $H\parallel[111]$. Filled circles in the map are peak positions of $C_P(T, H)$. (c,d) CMC results of specific heat $C(T, H)$ for (c) $(\delta, q) = (0, 0.85)$ with $J_{\rm nn} = 1.00$~K and for (d) $(\delta, q) = (-0.54, 0.5)$ with $J_{\rm nn} = 0.92$~K. The values of $J_{\rm nn}$ have been determined from the comparison of $\chi_{\rm Gnd}$ with the QMC results for each case. Solid, dashed, and dotted lines in (b), (c), and (d) are guides to the eyes. {\color[rgb]{0,0,0} Labels in maps (b) and (c) indicate assigned states from the analysis with Eq.(\ref{eq.1}) and CMC simulations~\cite{Kadowaki_MC}; i.e., (I) a paramagnetic paraquadrupole state, (II) the 3D PAF state, and (III) the 2D PAF state.} }

132 — 1506.04562

\caption{(a) Average swimming speed $\bar{v}$ (left axis) and (b) non-motile fraction $\beta$ for selected \ecoli\, coli strains, from one batch culture in each case, before washing (in TB, white) and after washing (in BMB+1.5~mM glucose, grey or in BMB, black) at a cell density of$5\times 10^8$cells/ml. Red bars in (a) are diameters of solid, circular colonies in 22-hour LB motility plates (right axis). Note that for HCB437, the cells did not penetrate the agar but stayed on the surface.}

\caption{Simultaneous measurements of average swimming speed $\bar{v}$ ({\color{red} $\circ$}, left axis) and [\ce{O2}] ($\square$, right axis), using fluorescence lifetime imaging microscopy (FLIM) vs. time in a sealed capillary for WT AB1157 at $\approx 10^9$~cells/ml and [\ce{Glu}]=0.5~mM. For FLIM, $\SI{47}{\milli\mbox{M}}$ of the \ce{O2}-sensitive dye ruthenium tris(2,2'- dipyridyl) dichloride hydrate (RTDP) was added to BMB (which did not affect the cells' motility). RTDP was excited using sub-ps pulses ($\lambda=450$nm, $\sim$1mW at 1MHz repetition rate) and its fluorescence imaged using a gated intensified CCD camera (Picostar HR-12QE, LaVision GmbH, Germany). The data was fitted with a single exponential decay yielding a (homogeneous) fluorescence lifetime map, which was averaged to estimate [\ce{O2}]. Inset: $\bar{v}$ at $5\times 10^8$~cells/ml to better show the saturation before \ce{O2} depletion. Data sets are from independent batch cultures with $\bar{v}(t=0)\approx \SI{12}{\micro\meter\per\second}$ and $\approx \SI{15}{\micro\meter\per\second}$, respectively. }

133 — 1506.04788

\caption{(Color online) Distributions of the 3--tangle ({\color{blue} +}) and its square (${\color{blue} \triangledown}$) for three-qubit pure random states. The solid lines are approximations with Beta distributions (\ref{dtau}) at left, and (\ref{dtau2}) at right.}

\caption{(Color online) Distribution of absolute value of the hyperdeterminant $P(T)$ ({\color{blue} +}) for four qubit pure random states in a log-log plot.}

\caption{(Color online) The mean of the maximum component (left) and the geometric measure of entanglement (right) for random states of a tri--partite system as a function of the qudit size. Bullets correspond to numeric simulations: ({\color{red} $\square$}) stands for the greatest tensor element $\lambda_{\max}$; ($\triangledown$) the maximum component of the HOSVD co-tensor $\lambda_H$; ({\color{blue} $\circ$}) stands for PARAFAC overlap $\lambda_P$ and ({\color{greenm} $+$}) refers to the overlap with the closest separable state maximized by LU. The solid red line ({\color{red} \bf ---}) is the result (\ref{maxelem}) with $N=d^3$, the solid lines ({\color{black} \bf ---}) and ({\color{blue} \bf ---}) are the best linear fits for HOSVD and PARAFAC, respectively}

134 — 1506.04808

\caption{The posterior distribution of the observational constraints on the comoving temperature of inflaton particles (the temperature axis is in units of ${\rm Mpc}^{-1}$) obtained from COSMOMC using Planck 2013 and WMAP nine year data. The $T$ only case refers to the analysis without taking into account the modification of the inflaton mode functions due to the preceding radiation era (akin to Ref. \cite{2006PhRvL..96l1302B}) while the $T$ and $\delta N$ case refers to the analysis including the modified mode functions, as in Ref. \cite{Das:2014ffa}. \textcolor{green}{The probability is normalised to the maximum probability.}}

135 — 1506.05228

\caption{\red{Comparison of the constraints on the amplitude of the delta-function power spectrum eq.(\ref{delta}), obtained from PBH, BBN, CMB, \blue{PTAs,} LIGO and Virgo. For more details see \cite{prep}. \cyann{\umi{The} PBH constraints in this figure are obtained assuming Gaussianity of primordial tensor perturbations and calculating the PDF of induced radiation perturbations numerically in \cite{prep}. The thick and dashed blue lines correspond to CMB constraints for adiabatic and homogeneous initial conditions of GWs, respectively. Note that the constraint from PTA is more than two orders of magnitude tighter than those from PBHs at $k_p\sim 4\times 10^6\mathrm{Mpc}^{-1}$. The frequency corresponding to each $k_p$, $f_p=ck_p/2\pi$, is also shown. See the text for more details. } } }

136 — 1506.05292

\caption{Normalized imaginary part of the susceptibility for the magnetic resonances labeled as \textcolor[rgb]{1,0,0}{1} and \textcolor[rgb]{1,0,0}{2} respectively. Resonance \textcolor[rgb]{1,0,0}{1} shows a localized character (quasi uniform mode) with the strongest excitation in the center region of the stripe. Resonance \textcolor[rgb]{1,0,0}{2} exhibits a harmonic dependence across the width of the stripe as expected for a standing spinwave with a wavelength of $734 \nm$. The profile along the dashed line is shown in fig. \ref{fig5}.}

137 — 1506.05577

\caption{Resulting background fit of the star KIC~12008916, as derived by \diamonds. The original PSD is shown in gray, while a smoothed version with boxcar width set to $\Delta\nu/5$, with $\Delta\nu$ taken from \cite{Huber11}, is shown as a black line to guide the eye. The red thick line represents the background model without the Gaussian envelope, while the red dotted line accounts for the additional Gaussian component. The individual components of the background model are shown by blue dot-dashed lines.}

138 — 1506.06135

\caption{\label{fig:lensing_signal} Halo mass consistency and assembly bias from the WL signal.: The data points with errorbars in the top and middle panels show the excess surface mass density profile as a function of the cluster-centric projected radius (in comoving units), obtained from the WL measurements for the large- and small-$\ave{R_{\rm mem}}$ subsamples of \redms clusters (see Eq.~\ref{eq:aveR}), respectively. The points from the top panel are reproduced in semi-transparent color in the middle panel for comparison. The mass profiles at small radii $R\simlt 10~h^{-1}{\rm Mpc}$ appear to have similar amplitudes with slightly different shapes, but show a difference in amplitude at $R\simgt 10~h^{-1}{\rm Mpc}$, as expected from {\it assembly bias}. The bold solid line shows the best-fit halo model, the thin solid line is the centered 1-halo term, the dashed line is the off-centered 1-halo term, while the dotted line corresponds to the stellar mass contribution from the central galaxy. Comparison between the dot-dashed lines in the two panels implies that the 2-halo term contributions, which arise from the average mass distribution surrounding the clusters, are different by a factor of 1.6. The bottom panel shows the ratios of the lensing signals, highlighting a clear deviation from unity at $R\simgt 15~$Mpc$/h$.}

139 — 1506.06226

\caption{\label{pic:dilution}(Color online) (a,d) The missing mass distribution, with the proton as the missing particle, and (b,e) the azimuthal angle difference of $\pi^0$ and proton for reaction (1), for a $\gamma p$ invariant mass of $W = 1.46$\,--\,1.48 (top) and 1.82\,--\,1.94\,GeV (bottom); butanol ({\scriptsize$\square$}), scaled carbon ({\color{red}$\circ$}), and the difference ({\color{blue}$\triangle$}). The distributions are shown after all other cuts discussed in the text are applied. From these distributions, the dilution factor (c,f) is determined. The gray band indicates the systematic uncertainty in the dilution factor due to uncertainties in the determination of the carbon scaling factor $s$. Since only events with all three particles detected in the calorimeter are considered an acceptance hole occurs for large cos($\theta$). The observed structures in the dilution factor are due to a combined effect of reduced efficiencies for clusters impinging onto detector boundaries and Fermi smearing.}

140 — 1506.06343

\caption{ \textcolor{blue}{An illustration of the pattern selection process (Sec.~\ref{subsubsec:pattern_Sselection}). For each pattern on the left, the image patches which form the corresponding mid-level visual elements are shown on the right. The red number underneath each patch is the image index. Since the top and bottom pattern cover $4$ and $5$ unique images, the coverage values of them are $4$ and $5$ respectively.}}

\caption{Pipeline to construct a Bag-of-Elements representation, \textcolor{blue}{which has been used in previous works as well~\citep{DBLP:conf/eccv/SinghGE12,DBLP:conf/nips/DoerschGE13, DBLP:conf/cvpr/JunejaVJZ13,bansal2015mid}.}}

\caption{\textcolor{blue}{Analysis of the transaction length on the VOC 2007 dataset using the \emph{VGG-VD} model. Other parameters are frozen.}}

\caption{\textcolor{blue}{Analysis of the merging threshold $Th$ in Algorithm~\ref{alg:merging} on the VOC 2007 dataset using the \emph{VGG-VD} model. Other parameters are frozen.}}

\caption{\textcolor{blue}{An illustration of the three firing types of mid-level elements. In the image, ground-truth object instances of the underlying category (\eg, ``person'') are overlaid in green while instances of other categories (\eg, ``cow'') are overlaid in red. Obviously, the firing (1) fires on the ground-truth object while firings (2) and (3) belong to object and scene context respectively.} }

141 — 1506.07081

\caption{Leading edge simulations for both mock- (left) and EGF- (right) treated cell populations. The red dots represent the experimental leading edge locations over time, and the solid green and dashed blue lines represent the leading edge locations for Models H and P, respectively. Note that while both models can match the mock data relatively well, model P appears to fit EGF data better than model H.\textcolor{blue}{{} }\textcolor{black}{The black dash-dot lines denote one standard deviation of the leading edge locations, and their computation is described in the appendix.} \label{fig:Leading-edge-simulations}}

142 — 1506.07515

\caption[]{Examples of elementary shapes shown with their characteristic frequency $\nu$. See {\purple https://www.youtube.com/watch?v=waXWOv0YqFE} for a movie showing how the shape varies with continuously changing frequency. }

143 — 1506.08138

\caption{$(a)$ Evolution of the jet half-width normalized with $D_e$ vs stream-wise direction $x^*$. Symbols: ($\blacksquare$): $r_{1/2}^s$ for $J_S$; ({\color{black} $\circ$}): $r_{1/2}^s$ for $J_F$; ({\color{black} $\blacktriangle$}): $r_{1/2}^s$ for $J_6$; ({\color{black} $\times$}): $r_{1/2}$ for $J_C$. Dashed lines for the $r_{1/2}^l$. $(b)$ Evolution of $\left\langle \Omega_{\theta} \right\rangle$ and $\left\langle\Omega_x \right\rangle$ vs the stream-wise direction $x^*$. Symbols: ($\square$): $\left\langle \Omega_x \right\rangle$ for $J_S$; ($\blacksquare$): $\left\langle\Omega_{\theta} \right\rangle$ for $J_S$; ({\color{black} $\circ$}): $\left\langle \Omega_x \right\rangle$ for $J_F$; ({\color{black} $\bullet$}): $\left\langle\Omega_{\theta} \right\rangle$ for $J_F$; ({\color{black} $\Diamond$}): $\left\langle \Omega_x \right\rangle$ for $J_6$;({\color{black} $\Diamondblack$}): $\left\langle\Omega_{\theta} \right\rangle$ for $J_6$; ({\color{black} $\triangle$}): $\left\langle \Omega_x \right\rangle$ for $J_C$; ({\color{black} $\blacktriangle$}): $\left\langle\Omega_{\theta} \right\rangle$ for $J_C$.}

144 — 1506.08377

\caption{Summary of the sensitivities obtained for $\sin^2 \theta_W$ (1$\sigma$) and for the EM neutrino parameters (90\% C.L.) at the TEXONO experiment. \red{The results refer to various sensitivities and quenching factors.} Comparing with Ref.~\cite{Kosmas:2015sqa} one sees that a substantial improvement in the sensitivity for the weak mixing angle $\sin^2 \theta_W$, the magnetic moment $\mu_{\bar{\nu}_e}$ parameter and the neutrino charge-radius $\langle r_{\bar{\nu}_e}^2 \rangle$ w.r.t. the COHERENT proposal.}

145 — 1506.08765

\caption{Clustering phenomenon. By enlarging the figure, the reader will distinguish one cluster of black triangles ({\footnotesize \color{black} $\triangledown$}) and one of cyan squares ({\footnotesize \color{cyan} $\square$}) near the origin, which correspond to the locations obtained by Brand et al.~and Govindu respectively.}

146 — 1506.08800

\caption{Workflow for \redis and \tool}

\caption{\bench for \redis vs \tool (times in seconds, median of 11 trials)}

\caption{\bench with single instructions for \redis vs \tool (times in seconds, median of 11 trials)}

\caption{\bench with 10 pipelined instructions for \redis vs \tool (times in seconds, median of 11 trials)}

147 — 1506.09041

\caption{Rheology in the colloidal, intermediate and granular size regimes. (a) Relative viscosity $\eta$ versus shear stress in units of Pa and $k_BT/d^3$ for $d=404$ nm spheres. At $\sigma <10$~Pa we performed downward sweeps in $\gdot$ and at $\sigma >10$~Pa, upward sweeps in $\sigma$, at volume fractions (\%) as labelled. Solid lines: fits based on~\cite{Wyart2014}; finely-dotted lines: schematics based on literature data (with sparsely-dotted = unstable states). Samples shear thicken above a $\phi$-independent onset stress $\sigma^*$ (vertical dashed line). Colour scheme: blue = frictionless interactions ($\sigma<\sigma^*$), black = shear thickening, red = frictional interactions ($\sigma \gg \sigma^*$). The unshaded region is accessible using our rheometer, which reaches maximum and minimum shear rates of $8000$~\rs and $10^{-3}$~\rs\! respectively, and a minimum stress of$10^{-2}$~Pa; the maximum accessible stress is set by a $d-$dependent fracture stress, $\sigma^\dagger$. % (b) Main: $\eta(\phi)$ for the the limiting high-shear viscosity, $\eta_1$, in (a) (\textcolor{blue}{$\blacksquare$}), and the lower, $\eta_1(\phi)$ (\textcolor{red}{$\filledmedsquare$}), and upper, $\eta_2(\phi)$ (\textcolor{red}{$\medsquare$}), branches in (c). Solid red line: least squares fit to $\eta_2(\phi) = A(1-\phi/\phi_m)^{-n}$ with $A=0.20(9)$, $\phi_m = 0.558(5)$ and $n=2.2(2)$. $\eta_1(\phi)$ data for other sizes of PMMA spheres in this work: $d = 912$~nm ($\medtriangledown$) and 1800~nm ($\filledmedtriangledown$). Other symbols: literature high shear viscosities (with $\phi$ shifted by up to $5$\%) for sterically stabilised PMMA \cite{Phan1996, Petekidis2004, Frith1996, D'Haene1993}, sterically stabilised silica \cite{VanderWerff1989} and glass beads \cite{Brown2013}; ($\bullet$) and ($\medcircle$) = lower and upper branches from \cite{D'Haene1993}. Inset: $\eta_2(\phi)$ versus ($\phi_m - \phi$) including the upper branch from \cite{D'Haene1993}. % (c) $\eta(\sigma)$ for $d = 3770$~nm spheres, all from upward $\sigma$ sweeps apart from $\phi = 58$\%, which is from a $\sigma$ downward sweep. The flow in both (a) and (c) was unsteady for $\phi \geq 0.56$, and points represent temporal averages. % (d) Schematic for the rheology of $d=350$~\micron\; spheres calculated from theoretical fits to\cite{Wyart2014} in (c). }

148 — 1507.00439

\caption{Excess intrusion and extrusion pressures with respect to pure water measured for the electrolytes mentioned in the legend. The x-scale is the van 't Hoff osmotic pressure $\Pi=icRT$ with $i$ the number of ions per salt molecule and $c$ the salt concentration. The values of our own data ($\blacksquare$) are listed in supplementary Tables 1 to 5 \cite{sup}. ({\Large $\diamond$}) are data from Tzanis et al \cite{Tzanis2014}, ({\large$\pentagon$}) are data from Ortiz et al \cite{Ortiz2014}. The straight line ($\cdots$) corresponds to the van 't Hoff law of osmotic pressure. The continuous line (\textcolor{Fuchsia}{$\relbar$}) is a simulation of the osmotic pressure of NaCl solutions from Luo et al \cite{Luo2010}. The inset is a magnification of our data with ($\Circle$) intrusion pressure and ($\bigtriangleup$) extrusion pressure.}

149 — 1507.00598

\caption{Summarization of physical-layer attacks in different stages of the CR cycle.}{\includegraphics[scale=0.65]{Tab1.eps}\label{Tab1}}

150 — 1507.01055

\caption{ Phase diagram obtained from simulations of Eq.~\eqref{e:dimless} for \markblue{one particular set of random initial conditions} showing the emergence of turbulent nematic states for supercritical active self-advection. (\textbf{a})~We observe convergence to defect-free stripes (blue, panel~\textbf{c}, Movie~S3), time-periodic and stationary defect lattice solutions (green, panel~\textbf{d}, Movie~S4), periodic defect creation and annihilation events (black, Movie S6), and chaotic dynamics (red, panel~\textbf{b}, Movie~S1). The white line indicates the analytical estimate ~\mbox{$D_c=2(-\gamma_2+\sqrt{\gamma_2^2+2})$} for the transitions between ordered and chaotic states. (\textbf{b}-\textbf{d}) Examples of the states identified in \textbf{a} with $-\frac{1}{2}$-defects (black) and $+\frac{1}{2}$--defects (white). Panel \textbf{d} highlights the antipolar ordering of $+\frac{1}{2}$-defect orientations (red bars); \markblue{see also Movie~S5 for a realization of this state in a larger simulation domain}. \label{fig:phase} }

\caption{\markblue{Strong and weak antipolar ordering of $+\f{1}{2}$-defects as (\textbf{a},\textbf{b}) observed in experiments~\cite{2015DeCamp} and (\textbf{c},\textbf{d}) predicted by our theory based on 2D simulations with periodic boundary conditions. Light-blue markers: $-\f{1}{2}$-defects. Yellow markers: $+\f{1}{2}$-defects. Red bars: orientation of $+\f{1}{2}$-defects. (\textbf{a})~$+\f{1}{2}$-defects in thin ALC films (thickness $h\sim 250\,$nm) show strong nematic alignment. (\textbf{b})~$+\f{1}{2}$-defects in thicker ALC films ($h\sim 1\,\mu$m) are more disordered. (\textbf{c},\textbf{d}) For weak effective hydrodynamic coupling $D$, simulations show antipolar ordering, which is inhibited for larger values of $D$. The average number of defects in the full simulation box is approximately (\textbf{c}) 240 and (\textbf{d}) 350. Figures \textbf{a} and \textbf{b} kindly provided by S. DeCamp and Z. Dogic.} \label{fig:nematic_1} }

\caption{ \markblue{Increasing activity and film thickness decrease antipolar ordering in simulations. (\textbf{a},\textbf{b}) Maxima of the numerically obtained local pair orientation PDFs $p(\theta_{ij}|r)$ signal antipolar local ordering of $+\f{1}{2}$-defects as they are separated by the typical defect-lattice spacing. The defect distance $r$ is specified in units of the mean nearest-neighbor distance $r_0$ between $+\f{1}{2}$-defects. (\textbf{c},\textbf{d})~Polar $P(r)$ and nematic $S(r)$ correlation functions for $D=1.5$ (red) and $D=3$ (blue). Increasing the effective hydrodynamic coupling~$D$ leads to stronger mixing and hence decreases nematic order. The simulation parameters correspond to those given in Fig.~\ref{fig:nematic_1}\textbf{c},\textbf{d}. \label{fig:nematic_2}} }

151 — 1507.02144

\caption{{\fontsize{\captionfonotsize pt}{0pt}\selectfont\color{bmv@captioncolor} AP for animal classes of VOC 2007, comparing DPM with GDPM with different number of parts. Sub (super) indices indicate the number of positive (negative) parts.}}

\caption{{\fontsize{\captionfonotsize pt}{0pt}\selectfont\color{bmv@captioncolor}Analysis of false positives of \textit{cow} using \cite{HoiemCD12}. $DPM$ (left), $GDPM^{2}_6$ (middle) and $GDPM^{1}_{7}$ (right) with similar part initialization. The volumes show the commulative number of top detections within 5 different categories: true detection (white), localization FP (blue), background FP (purple), similar objects (red), other objects (green). Red (dashed red) line is the percentage of recall at each number of detections using 50\% (10\%) overlap threshold. GDPM with careful initialization (right figure) slightly reduces confusion with similar objects.}}

\caption{{\fontsize{\captionfonotsize pt}{0pt}\selectfont\color{bmv@captioncolor} Example of an LVM scoring function on three latent variables (left), a Simple GLVM scoring function (middle), and a GLVM in its canonical form. The models become more and more complex from left to right.}}

152 — 1507.02805

\caption{\kemperecolor}

153 — 1507.03148

\caption{Our proposed head pose based cascaded face alignment procedure (path in \textcolor{cyan}{cyan} color) vs. conventional cascaded face alignment procedure (path in \textcolor{red}{red} color). }

154 — 1507.03167

\caption{Structure of the matrix $\langle n | \hat{P}^{(c)}_{q p} | n' \rangle$ of Eqs. (\ref{nPn'}), (\ref{nPn' 2}), for $N=5$ and $c=0$. The non-zero, off-diagonal matrix elements for fixed $q$ (which obey $n+n'=2q+1$) appear on the various secondary diagonals. For a given $q$, the diagonal matrix element $\delta_{nq}\delta_{n'q}$ {\color{blue}(i.e., $n=n'=q$)} does {\em not} lie on the same secondary diagonal as the off-diagonal matrix elements; this is emphasized for the particular case $q=1$, where the various matrix elements have been highlighted and and a box has been drawn around each one of them. }

155 — 1507.03370

\caption{Calculated phase $\phi+\tilde \phi$ as a function of the wavelength for two different lengths of the compensation crystal (constant offset subtracted). A flat phase is obtained for $\tilde L=154$\,mm (\blacksolid) for the two target wavelengths $\lambda_s=894.3$\,nm and$\lambda_i=1313.1$\,nm (vertical lines). The plateaus of the flat phases are shifted considerably for$\tilde L = 153$\,mm (\blackdashed). \label{fig:yvo4length}}

156 — 1507.03611

\caption{(a) HD 30495 combined $S$-index time series, including data from MWO ({\color{Gray}$\bullet$}), SSS ($\circ$), SMARTS ({\color{Cyan}$\blacksquare$}), CPS ({\color{ForestGreen}$\blacktriangle$}), and HARPS ({\color{Yellow}$\blacktriangledown$}), along with seasonal means ($\bullet$), with error bars representing the error of the mean. The red curve is a 3-component sine wave model of the stellar cycles, \revone{while the horizontal red line is our reference point for global activity minimum.} (b) Zoomed portion highlighting higher-cadence SMARTS data, with a $P = 1.58$ yr sine wave plotted for comparison (blue dashed curve). (c) APT differential photometry brightness measurements in the combined Str{\"o}mgren $b$ and $y$ bands, in milli-magnitudes. Differences shown are HD 30495 nightly measurements ({\color{YellowGreen}$\bullet$}) and seasonal means ($\bullet$) with respect to the comparison stars, as well as the difference between the two comparison stars ($\square$) (d) Seasonal mean differential brightness difference in the $b$ and $y$ bands, in milli-magnitudes, with colored regions indicating brighter $b$ (blue) and $y$ (green) emission. A horizontal dotted line indicates the grand mean in all panels. \revone{Magnitude scales are inverted such that brightness increases in the upward direction. The observations shown in this figure are available in the electronic version of this publication. See tables \ref{tab:obs_S} and \ref{tab:obs_by} in the Appendix.}}

157 — 1507.03753

\caption[Out-of-phase NNM]{Out-of-phase NNM;% \includegraphics{img/firstNNMLeg1.pdf}% FE method,% \includegraphics{img/firstNNMLeg2.pdf}% Taylor series; (a) $x_1,y_1,x_2$; (b) $x_1,y_1,y_2$}

\caption[In-phase NNM for different values of $k_b$]{In-phase NNM for different values of $k_b$; (a)-(b) $k_b=4.7$, (c)-(d) $k_b=4.5$, (e)-(f) $k_b=4.1$;% \includegraphics{img/invManK2Leg1.pdf}% FE method,% \includegraphics{img/invManK2Leg2.pdf}% Taylor series}

\caption[Comparison of time series]{Comparison of time series;% \includegraphics{img/TSleg1.pdf}% Numerical integration,% \includegraphics{img/TSleg2.pdf}% Taylor approximation; initial condition: $x_1(0)=-0.23$, $x_2(0)=-0.76$, $y_1(0)=-0.4$, $y_2(0)=-0.86$; (a) $x_1$, (b) $x_2$, (c) $y_1$, (d) $y_2$}

\caption[Decomposition of trajectories in different terms]{Decomposition of trajectories in different terms;% \includegraphics{img/MDleg1.pdf}% $1^{\text{st}}$ order term,% \includegraphics{img/MDleg2.pdf}% $3^{rd}$ order term,% \includegraphics{img/MDleg3.pdf}% $5^{th}$ order term; % (a) $k_b=4.7$ ($x_1(0)=-0.57$, $x_2(0)=-0.55$, $y_1(0)=-0.44$, $y_2(0)=-0.66$); % (b) $k_b=4.1$ ($x_1(0)=-0.23$, $x_2(0)=-0.76$, $y_1(0)=-0.4$, $y_2(0)=-0.86$)}

\caption[Comparison of $5^{\text{th}}$ order approximation and numerical integration]{Comparison of $5^{\text{th}}$ order approximation and numerical integration;% \includegraphics{img/LOAleg1.pdf}% Numerical integration,% \includegraphics{img/LOAleg2.pdf}% $5^{\text{th}}$ order approximation; % (a) $k_b=4.7$ ($x_1(0)=-0.57$, $x_2(0)=-0.55$, $y_1(0)=-0.44$, $y_2(0)=-0.66$), % (b) $k_b=4.1$ ($x_1(0)=-0.23$, $x_2(0)=-0.76$, $y_1(0)=-0.4$, $y_2(0)=-0.86$)}

158 — 1507.03822

\caption{ {\color{red}The mixing coefficients ${\cal N}_{\ell N}$ ($\ell = e, \mu, \tau$) for the total decay width $\Gamma_N$, Eqs.~(\ref{GN1})--(\ref{calK}), as a function of the mass $M_N$ of the neutrino $N$. The left-hand figure is for the case of Majorana neutrino \cite{CKZ2}; the right-hand figure for the case of Dirac neutrino \cite{CDKZ}.} }

\caption{ {\color{red}Lepton flavor conserving decays $B^{-} \to \tau^{-} N \to \tau^{-} {\overline \nu}_{\ell} M^{(0)}$ and $B^{-} \to \tau^{-} N \to \tau^{-} \ell^{+} M^{-}$, where $M$ is a pesudoscalar or vector meson.} }

\caption{{\color{red} Lepton flavor violating decays $B^{-} \to \tau^{-} \ell^{-} M^{+}$ mediated by a Majorana neutrino $N$.} }

159 — 1507.03912

\caption{(a) Comparison between the self-similar shape of the sheet predicted by the time-dependent PDE (\ref{eq:NDtf}) and the first thirteen solutions $H_m(\eta); m\in [1-13]$ of the similarity ODE (\ref{eq:sim}-7). We note that only the first fundamental solution $H_1(\eta)$ of (6-7) is consistent with that of the PDE (\ref{eq:NDtf}) ({\color{red}$\circ$}). (b) Numerical simulations of the PDE (\ref{eq:NDtf}) using the initial condition associated with the self-similar shape $h_0=h(x,t=0)=H_3(\eta)$ shows the evolution of the shape of the sheet for four different time points $t=[0,~0.28,~0.336,~0.341]$. We find that the numerically measured contact time is $t_C=0.3412$ and the contact points are $x_C=[-1.48,1.48]$. (c) By using the rescaled variables $H(\eta)=h(x,t)/(t_C-t)^{\frac{1}{3}}$,$\eta=(x-x_C)/(t_C-t)^{\frac{1}{3}}$ with $x_C=1.48$ corresponding to the contact on the right side, we see that the shapes of the sheet obtained in Fig. 2a collapse onto the universal shape of the fundamental solution of (6-7) $H_1(\eta)$ (shown in red). The deviations away from the contact point are due to the effects of the second touchdown at $x_c=-1.48$. (d) To distinguish the different discrete solutions $H_m(\eta)$ satisfying (6-7) we plot the scaled bending pressure $H_m''''(\eta)$, and see that the number of maxima/minima scales with $m$. Thus, for $m>1$, the wavelength of the pressure oscillations decreases as $m$ increases; these solutions are therefore strongly damped out in comparison with the fundamental solution of $H_1(\eta)$ (6-7), which alone survives as we approach touchdown. \label{fig:groupedfig2}}

160 — 1507.04003

\caption{Schematic of the tree network, constructed by arranging operators according to their size. The identity operator is on the extreme left, followed by the single-site operators, and so on. The total number of operators is $D^{2L}$, while the fraction in the $n^{th}$ layer is $f_{n}=\frac{(D^{2}-1)^{n}}{D^{2L}}\left(\protect\begin{array}{c} L\protect\\ n \protect\end{array}\right)$, which is maximum for $n=(1-1/D^{2})L=\frac{3}{4}L\equiv n^{*}$ for $D=2$. Note that $f_{n}$ grows exponentially with $n$ for small $n$, so the apparent linear growth in the number of dots with $n$ in the figure is for ease of depiction and should not be taken literally.\label{fig:tree-schematic}} \end{figure} The orthogonality of the operator basis allows us to expand an arbitrary many-body Hermitian operator $A(t)$ in this basis as $A(t)=\sum_{n,\ell}\psi_{n\ell}(A;t)O_{\ell}^{(n)}$, with real coefficients: \begin{equation} \psi_{n\ell}(A;t)=\mbox{Tr}\left(A(t)O_{\ell}^{(n)}\right)\label{eq:wavefunction-def} \end{equation} The vector $\psi_{n\ell}(A;t)$ can be viewed as a "single-particle wavefunction" of a particle hopping on the graph we defined. There are two key requirements for interpreting $\psi_{n\ell}(A;t)$ as a sensible single-particle wavefunction within first quantization. Firstly, the total probability density of the particle must be conserved; this is guaranteed because $\left\{O_{\ell}^{(n)}\right\}$ form an orthonormal basis for operators in the Hilbert space, which implies \begin{equation} \sum_{n,\ell}\left|\psi_{n\ell}(A;t)\right|^{2}=\mbox{Tr}\left(A^2\right)\label{eq:psi-norm} \end{equation} a manifestly invariant quantity under unitary time-evolution of $A$. Secondly, it must satisfy Schrodinger's equation. Indeed, $A(t)$ follows Heisenberg time evolution: $\dot{A}(t)=i[H,A(t)]$, so the time-evolution of $\psi_{n\ell}$ is given by \begin{eqnarray} i\frac{\partial}{\partial t}\psi_{n\ell}(A;t) & = & \mbox{Tr}\left(\left[A(t),H\right]O_{\ell}^{(n)}\right)\nonumber \\ & = & \mathcal{H}_{\ell\ell'}^{nn'}\psi_{n'\ell'}(A;t)\label{eq:Schrodinger} \end{eqnarray} where the hopping matrix element between two nodes $O_{\ell}^{(n)}$ and $O_{\ell'}^{(n')}$ is defined as \begin{equation} \mathcal{H}_{\ell\ell'}^{nn'}=\mbox{Tr}\left(\left[H,O_{\ell}^{(n)}\right]O_{\ell'}^{(n')}\right)=-\mathcal{H}_{\ell'\ell}^{n'n}=-\left(\mathcal{H}_{\ell\ell'}^{nn'}\right)^{*}\label{eq:hopping} \end{equation} Thus, $\psi_{n\ell}(A;t)$ is a sensible wavefunction, and its dynamics are governed by the hopping Hamiltonian (\ref{eq:hopping}). Crucially, if the physical Hamiltonian contains at most $k$-spin terms, $\mathcal{H}_{\ell\ell'}^{nn'}$ vanishes for $|n-n'|\ge k$ and therefore, satisfies a notion of locality on the graph. Thus, the time-evolution of a general operator in a many body system has been recast into the problem of a single particle governed by a local hopping Hamiltonian on a high-dimensional graph. The ``energies'' of the particle consists of all possible differences $E_{i}-E_{j};i,j=1\dots D^{L}$ between pairs of energies of $H$, so the single-particle spectrum that results from a generic $H$ with no degeneracies is particle-hole symmetric with $D^{L}$ zero eigenvalues and $D^{2L}-D^{L}$ non-zero ones. The above construction is reminiscent of mappings of states in Fock space to a Cayley tree, which allows one to view integrable systems with local conservation laws in real \cite{Altshuler1997} and momentum \cite{Neuenhahn2012} space as a localized particle on a suitably defined tree. Our construction, on the other hand, describes the Fock space of operators instead of that of states, and thus is closer to the approach adopted by Ref. \cite{Ros2015} for constructing integrals of motion to the describe the many-body localized phase. The graph construction facilitates extracting different kinds of information about quantum ergodicity with different choices of $A$. If we choose $A$ to be the density matrix $\rho$ for an eigenstate that satisfies the ETH, then $\psi_{n\ell}(A)$ is expected to coincide with its thermal value, determined only by the energy density in that state, for small $n$, but can be depend on other details for larger $n$. On the other hand, picking $A$ to be the difference between two density matrices tells us how various correlators differ in the two states, and what kind of observables must be measured in order to distinguish between them. Finally, if $A$ is a physical observable $O$, its time-evolution tells us how this physical quantity evolves into a superposition of other operators with time. In particular, suppose we start from a small-$n$ operator and let it evolve in time. In general, it will evolve into a superposition containing many large-$n$ operators. Equivalently, one can think of $\psi_{n\ell}(O;t)$ as an infinite temperature correlation function, so its time evolution tells us how small operators develop correlations with large operators over time. In the language of the hopping particle, this corresponds to the particle starting near the low-$n$ end of the graph and spreading towards larger $n$ nodes. Thus, chaotic behavior of operators turns into delocalization of the particle on the graph. \section{$n$-weight and $n$-distinguishability\label{sec:PnThetan}} A central quantity that we will work with in this paper is the $n$-weight of an operator $A$, $P_{n}(A)$, defined as \begin{equation} P_{n}(A)=\sum_{\ell}\left|\psi_{n\ell}(A)\right|^{2}=\sum_{\{r_{\alpha},i_{\alpha}\}}\left|\mbox{Tr}\left(AO_{\{r_{\alpha},i_{\alpha}|\alpha=1\dots n}\right)\right|^{2}\label{eq:Pn-def} \end{equation} where the sum over $\ell$ runs over all possible choices of $n$-site operators. In the graph picture of Sec. \ref{sec:fock-space}, this is the total probability density in the $n^{th}$ layer of the graph. Physically, $P_{n}(A)$ tells us how complex the correlators one must measure in order to reconstruct $A$ are. Computing it directly is computationally taxing, as it entails computing $D^{2L}$ traces, one for each operator in the Hilbert space. Fortunately, the computation can be simplified via a generating function, as follows. Performing the sums over $\{i_{\alpha}\}$ for fixed $\{r_{\alpha}\}$ and using Eq. (\ref{eq:single-site-ortho}) gives \begin{equation} P_{n}(A)=\frac{1}{D^{L-n}}\sum_{R_{n}}\prod_{r\in R_{n}}\mbox{Tr}\left[(A\otimes A)W_{r}\right]\label{eq:PnW} \end{equation} where $R_{n}$ is a region of size $n$ (not necessarily connected), and the sum $\sum_{R_{n}}$ is over all $n$-site regions. Now we define a generating function \begin{eqnarray} F(A;z)=\sum_{n=0}^{L}z^{n}P_{n}(A) \end{eqnarray} which can be explicitly written as \begin{eqnarray} F(A;z) & = & \frac{1}{D^{L}}\mbox{Tr}\left[A\otimes A\prod_{r=1}^{L}\left(\mathbb{1}_{r}+DzW_{r}\right)\right]\label{eq:gen-func}\\ & = & \frac{1}{D^{L}}\sum_{R}(1-z)^{L-k_{R}}(Dz)^{k_{R}}\mbox{Tr}_{R}\left(\mbox{Tr}_{\bar{R}}A\right)^{2}\nonumber \end{eqnarray} Here the sum $\sum_{R}$ is over all regions $R$, composed of $k_{R}$ sites and $\bar{R}$ denotes the complement of $R$. $P_{n}(A)$ is then determined by Fourier transforming $F(A;e^{in\theta})$: \begin{equation} P_{n}(A)=\frac{1}{2\pi}\int\mathrm{d}\theta e^{-in\theta}F(A;e^{in\theta}) \end{equation} Since $n\in[0,L]$ is linear in system size, it is actually sufficient to calculate $F(A;e^{in\theta})$ for the $L+1$ discrete values of $\theta=\frac{2\pi}{L+1}m,~m=0,1,..,L$. Therefore we have translated the calculation of $P_{n}(A)$ for all $A$ to $L+1$ operator trace computations in the doubled Hilbert space. Alternatively, we can also use the second line of Eq. (\ref{eq:gen-func}) and calculate $P_{n}$ by performing $D^{L}$ partial trace calculations on a single copy of the system. When $A$ is a density matrix $\rho$, the generating function also makes it transparent that $P_{n}(A)$ is related to the second Renyi entropy of a region $R$, $S_{R}=-\log{\rm Tr}\left(\rho_{R}\right)^{2}=-\log\mbox{Tr}\left[\rho\otimes\rho\prod_{r\in R}X_{r}\right]$. From (\ref{eq:PnW}), we have \begin{equation} P_{n}(\rho)=\frac{1}{D^{L-n}}\sum_{R_{n}}\sum_{R_{k}\subseteq R_{n}}e^{-S_{R_{k}}}\left(-\frac{1}{D}\right)^{n-k}\label{eq:Pn-explicit1} \end{equation} The sum over $R_{n}$ and $R_{k}$ can be combined into a single sum by introducing suitable combinatorial factors. Defining $e^{-\mathbb{S}_{k}}=\left\langle e^{-S_{R_{k}}}\right\rangle _{R_{k}}$, i.e., the average of $e^{-S_{R_{k}}}$ over all $k$-site regions, we get \begin{equation} P_{n}(\rho)=\frac{1}{D^{L-n}}\left(\begin{array}{c} L\\ n \end{equation}\end{array}\right)\sum_{k=0}^{n}e^{-\mathbb{S}_{k}}\left(-\frac{1}{D}\right)^{n-k}\left(\begin{array}{c} n\\ k \end{array}\right)\label{eq:Pn2Sn} \end{equation} Thus, there is a simple relationship between $P_{n}$, the ``single particle density on the graph'', and entanglement properties of the many body state. Based on the $n$-weight defined for each operator $A$, we define a second quantity, the $n$-distinguishability between two operators $A_{1}$ and $A_{2}$, as \begin{equation} \theta_{n}(A_{1},A_{2})=\cos^{-1}\frac{P_{n}(A_{1})+P_{n}(A_{2})-P_{n}(A_{1}-A_{2})}{2\sqrt{P_{n}(A_{1})P_{n}(A_{2})}}\label{eq:thetan-def} \end{equation} $\theta_{n}\left(A_{1},A_{2}\right)$ is simply the angle between the two vectors $\vec{\psi}_{n}(A_{1})=\psi_{n\boldsymbol{\ell}}(A_{1})={\rm Tr}\left(A_{1}O_{\boldsymbol{\ell}}^{(n)}\right)$ and $\vec{\psi}_{n}(A_{2})$ defined similarly, i.e. the two ``single particle wavefunctions" corresponding to $A_{1}$ and $A_{2}$, projected to the graph sites corresponding to size-$n$ operators. The angle $\theta_{n}$ thus measures how different the two operators are if only $n$-site operators are measured. A small $\theta_{n}$ implies that $A_{1}$ and $A_{2}$ look similar in all size-$n$ measurements, while a large $\theta_{n}\sim\frac{\pi}{2}$ means $A_{1}$ and $A_{2}$ can be easily distinguished by $n$-site operators. This is sketched in Fig. \ref{fig:thetan-sketch}, where we have chosen $A_{1,2}$ to be two density matrices $\rho_{1,2}$ in anticipation of the discussion in the next section. In Sec. \ref{sub:thetan}, we will use $\theta_{n}$ to distinguish between neighboring eigenstates and show that indeed, they appear similar for simple observables and different for complicated ones. \begin{figure} \begin{centering} \includegraphics[width=0.48\columnwidth]{rho-vectors-small-n}\includegraphics[width=0.48\columnwidth]{rho-vectors-large-n} \par\end{centering} \caption{Schematic illustration of the behavior of neighboring eigenstates $\rho_{1}$ and $\rho_{2}$ on being projected onto the space of small-$n$ (left) and large-$n$ (right) operators. $\rho_{i}^{(n)}$ is shorthand for the projections $\vec{\psi}_{n}(\rho_{i})$ defined in the text. The three directions together depict the full Hilbert space of operators, while the horizontal plane represents its projection onto the space of operators of size $n$. $\rho_{1}$ and $\rho_{2}$ are mutually orthogonal vectors in the full space. However, they appear nearly identical when projected onto simple operators as shown on the left, but look quite different for larger operators as shown on the right. The angle $\theta_{n}$ will be calculated numerically for the non-integrable Ising model in Sec. \ref{sec:eth-ising}.\label{fig:thetan-sketch}} \end{figure} \section{eigenstate thermalization in the Ising model\label{sec:eth-ising}} \begin{figure} \begin{centering} \includegraphics[width=0.49\columnwidth]{Pndiff-vs-n-no-fn}\includegraphics[width=0.49\columnwidth]{Pndiff-vs-n-no-fn-int} \par\end{centering} \begin{centering} \includegraphics[width=0.49\columnwidth]{Pndiff-vs-n}\includegraphics[width=0.49\columnwidth]{Pndiff-vs-n-int} \par\end{centering} \caption{$P_{n}(\Delta\rho)$ vs $n$ (above) and $P_{n}(\Delta\rho)/f_{n}$ vs $n$ (below) for $\sim200$ randomly chosen pairs of neighboring eigenstates for $L=14$ sites for the ergodic (left) and the integrable (right) Ising model. Here, $\Delta\rho = \rho_1-\rho_2$ is the difference between the density matrices of the two eigenstates. The color is proportional to the density of states, with blue (red) representing states in regions of the spectrum with low (high) density of states. The black dashed line marks $P_{n}(\Delta\rho)=2f_{n}$.\label{fig:Pn-delta-rho}} \end{figure} In this section we apply the new measures we define to study eigenstate thermalization in a prototypical non-integrable spin model, namely, the 1D Ising model with transverse and longitudinal fields, given by \begin{equation} H=\sum_{r}\left(J\sigma_{r}^{z}\sigma_{r+1}^{z}+h_{x}\sigma_{r}^{x}+h_{z}\sigma_{r}^{z}\right)+h_{z}\sigma_{1}^{z}\label{eq:Hamiltonian} \end{equation} $H$ is integrable if any one of $J$, $h_{x}$ and $h_{z}$ vanishes, but is non-integrable otherwise. We choose $J=0.5$, $h_{x}=-0.74$ and $h_{z}=0.35$ as the non-integrable parameters, and $J=0.5$, $h_{x}=0.35$ and $h_{z}=0$ as the integrable ones. Open boundary conditions and the extra term on the first site, $h_{z}\sigma_{1}^{z}$, ensure that translation and inversion symmetries are broken so that there are no conserved quantities in the non-integrable case. This is unlike several recent works which retained translational symmetry and hence, conserved the total momentum \cite{Garrison2015,Kim2014,Rigol2009,Rigol2009a}. The energies and eigenstates are obtained by exact diagonalization of systems of upto $L=14$ sites. \subsection{Comparison of eigenstate $n$-weights\label{sub:Pn}} As stated in the introduction, the ETH says that the expectation values of simple operators are equal in nearby eigenstates of chaotic Hamiltonians, up to exponentially small corrections in the system size. This automatically ensures that each eigenstate resembles a ``microcanonical ensemble'', i.e., an equal admixture of nearby eigenstates, in the thermodynamic limit and hence yields the ETH as stated in the introduction. Thus, we first compare pairs of neighboring eigenstates $\rho_{1}$ and $\rho_{2}$ by computing the total squared difference in the expectation values of all operators of size $n$, \begin{equation} P_{n}(\Delta\rho)=\sum_{\ell}\left(\left\langle O_{\ell}^{(n)}\right\rangle_{\rho_{1}}-\left\langleO_{\ell}^{(n)}\right\rangle_{\rho_{2}}\right)^{2} \end{equation} where $\Delta\rho=\rho_{1}-\rho_{2}$, and study its dependence on $n$ and the energy of the pair. As shown in the upper panels of Fig. \ref{fig:Pn-delta-rho}, this quantity has the anticipated behavior for small $n$: it increases with $n$ and is larger when the density of states is lower. A closer inspection, however, reveals that the $n$-dependence seen here is deceptive, and cannot be used to declare eigenstate thermalization. In particular, the curves approximately follow the fraction of operators of size $n$, $f_{n}=\frac{(D^{2}-1)^{n}}{D^{2L}}\left(\begin{array}{c} L\\ n \end{array}\right)$ upto an overall proportionality constant. In fact, the ``infinite temperature'' eigenstates -- states near the middle of the spectrum where the density of states is highest -- have $P_{n}(\Delta\rho)\approx2f_{n}=\mbox{Tr}\left[(\Delta\rho)^{2}\right]f_{n}$. Moreover, the $n$-dependence is roughly the same even for the integrable Ising model. Thus, we conclude that the bare $n$-dependence of $P_{n}(\Delta\rho)$ is primarily determined by the number of operators of size $n$, not by the integrability properties of the Hamiltonian. Therefore we study the average density per site $P_{n}(\Delta\rho)/f_{n}$ -- the mean squared difference in the expectation values of size-$n$ operators between neighboring eigenstates, upto an overall proportionality constant of $D^{2L}$. As is shown in the lower panels of Fig. \ref{fig:Pn-delta-rho}, the $n$-dependence of this quantity is clearly different for ergodic and integrable systems. However, its $n$-dependence for the ergodic system is the exact opposite of what one would naively expect from ETH. Indeed, the lower panels of Fig. \ref{fig:Pn-delta-rho} show that on average, large operators are actually worse at distinguishing between neighboring eigenstates than small operators are, irrespective of whether the Hamiltonian is integrable or not. The fact that $P_{n}(\Delta\rho)/f_{n}$ decreases with $n$ for small $n$ simply says that on average, simple operators store more information about the state of the system compared to complicated ones. For integrable systems as well as for ground states of ergodic systems, this statement is easily understood because nearby eigenstates \emph{can }be distinguished by simple operators. Fig. \ref{fig:Pn-delta-rho} says that random simple operators can split finite energy density states of ergodic Hamiltonians as well, but the efficiency with which they can do so decreases with increasing density of states. For the infinite temperature states (i.e., states at the part of the spectrum with the largest density of states) $P_{n}/f_{n}$ is almost independent of $n$, which means that there is no difference between simple and complicated operators, since the state is essentially a random state in the Hilbert space. \begin{figure*} \begin{centering} \includegraphics[width=2\columnwidth]{Pndiff-vs-g-erg-and-int} \par\end{centering} \caption{Left: $\frac{1}{L}\log\left[g_{\infty}/g(E)\right]$ vs the energy density $E/L$ for various system sizes for the ergodic (above) and the integrable (below) Ising model. Except near the band edges, the curves are indistinguishable, indicating that $g(E)/g_{\infty}$ grows exponentially with $L$ with an energy density dependent exponent. Middle: $\log\left[P_{n}(\Delta\rho)/f_{n}\right]$ vs $\log\left[g_{\infty}/g(E)\right]$ for various $n$ for $L=14$ for the ergodic (above) and the integrable (below) Ising model, and straight line fits to the ergodic data. The fits are good for $n\ge3$. A similar fitting procedure for other system sizes yields the panels on the right, where we show that the slopes (above) and the intercepts (below) of the straight lines are simple functions of $n/L$. \label{fig:scaling}}

161 — 1507.04153

\caption{(color online): \tmg{Upper panel:} \tmg{$U/t$ dependence of} magnetization $(m_{\rm b},m_{\rm dw})$, and density of states $\mathcal{D}(E_{\rm F})$ \tmg{at the Fermi level}. %\tgr{Plot the density of states at zero for $U>5$} \tmg{Lower panel: $U/t$ dependence of} anomalous Hall conductivity of a pyroclore slab with a single magnetic domain wall, and domain-wall width $\lambda_{\rm dw}$ \red{in the unit of} %\sout{compared with} distances between nearest-neighbor Kagom${\rm \acute{e}}$ layers $a_{\rm K}$. \cyan{In comparison with the anomalous Hall conductivity with a single domain wall (shown as ``$\sigma_{\textcolor{black}{XY}}$ w dw"), we also show the anomalous Hall conductivity without domain walls shown as ``$\sigma_{\textcolor{black}{XY}}$ w/o dw."} \tmg{Three phases, degenerate helical metal, helical metal, and insulator are identified from the density of states.} %\red{w, w/o, $\lambda_{\rm dw}$} }

162 — 1507.04197

\caption[]{The conditions for a valid sequence of eigensteps for equal norm tight frames with $\mu=d$. A wedge% \tikz[baseline=-\the\dimexpr\fontdimen22\textfont2\relax]{ \node (b) at (1.5,0) {$\strut\lambda_{k,l}$}; \node[baseline=(b.base)] (a) at (0,0) {$\strut\lambda_{i,j}$}; \draw[fill, triangle path=1.5pt] (a) -- (b); }% denotes an inequality $\lambda_{i,j}\le\lambda_{k,l}$.}

163 — 1507.04305

\caption{ \textcolor{black}{\textit{Top left panel:} PV diagram ($PA$ = 286\degree) centred at the optical centre. The angular offset, in arcsec, is negative to the SE (left) and positive to the NW (right). \textit{Top right panel:} PV diagram \textcolor{black}{($PA$ =244\degree) along the line connecting knot E and D centred at knot E}. The angular offset, in arcsec, is negative to the NE (left) and positive to the SW (right). The horizontal arrows indicated the size of the HERA beam. \textit{Bottom \textcolor{black}{left} panel}: }\cg079 integrated \textit{Spitzer} IRAC 4.5 $\mu$m (\textcolor{black}{dashed line}) and \hto\(\textcolor{black}{solid line}) emission \textcolor{black}{in normalised units} along the $PA$ = \textcolor{black}{286}\degree\centred on the optical centre of the galaxy. The horizontal axis shows the offset from the optical centre in kpc, with positive values to the NW.\textcolor{black}{\textit{Bottom right panel: }SDSS g -- band image showing the orientation and lengths of the slices along which \textcolor{black}{the} PV diagrams where derived.} }

164 — 1507.04321

\caption{\textbf{Structure factors $S(q)$ and pair correlation function from reconstructed structure.} (\textbf{a}) The numerically simulated structure factor $S(q)$ (\solidrule) fits the experimental SAXS $S(q)$ (\dashedrule). The small-angle lamellar peaks' positions, from \textit{L. hesperus} (top) to \textit{A. gemmoides} (bottom), are correspondingly located at 0.83, 0.78, 0.60 and 0.61 nm\textsuperscript{-1} with errors of approximately $\pm 0.01$nm\textsuperscript{-1}. The low-$q$ region of the structure factor $S(q)$, i.e. the matrix knees, exhibit linearity in all cases and the slopes range from -2.1 to -2.7 on the log-log scale. (\textbf{b}) Pair correlation function $P(r)$ calculated from the reconstructed electron density maps. The correlation function $P(r)$ were calculated from a population of around 5000 crystals and the curves were numerically smoothed. The intermediate range crystalline ordering is reflected as the multiple correlation peaks observed on the $P(r)$ curves in the range of 7 to 40 nm. }

\caption{\textbf{Time evolution of the structure factor and the pair-correlation function $P(r)$ during the stochastic reconstruction on silk species \textit{A. gemmoides} .} (\textbf{a}) From bottom to top, as the stimulated annealing temperature $T$ drops, the calculated structure factor (\solidrule) converges to the experimental structure factor (\dashedrule), reducing the pseudo-energy $E$ as defined in Eq.\ref*{eq:engy}. (\textbf{b}) Initially at $E=0.4256$, the $P(r)$ function is absent from any correlation peak. As the stimulated annealing algorithm proceeded, the model build up intermediate range crystalline ordering which was reflected by the correlation peaks at 12, 15, 21, 26, 32, 37 and 42 nm on the top curve ($E=0.0021$) . The correlation function $P(r)$ is sampled from a model contains 4721 $\beta$-sheet crystals. }

165 — 1507.04532

\caption{ VAF at real frequency axis (a) and in time representation (b) for $T=1.4$ and $\rho=1$: results of MD simulation. The inset in (b) shows long time tail of $Z(t)$ in the double logarithmic scale; the red bullet \textcolor{red}{$\bullet$} points the time scale $\Delta t=2\pi/\Delta \omega$, where $\Delta \omega$ is the half-width of the branch cut. Graphs (c) and (d) show analytical continuation of $Z(\omega)$ into complex $\omega$-plain using multipoint Pade approximation built on top of $1400$ uniformly distributed knot-points in $(0,6.3\omega_{\rm max})$. 3D plot of $|Z(\omega)|$ in the complex $\omega$-plane; the black curve in (a) is $Z(\omega)$ at real $\omega$. Regular behaviour of $Z(\omega)$ for small imaginary frequencies ends abruptly by the ``walls'' of singularities. Graph (d) shows $\ln|Z(\omega)|$ in the complex frequency domain $\Real\omega,\,\Imag\omega\in(-4.5\omega_{\rm max},4.5\omega_{\rm max})$.}

\caption{ VAF at real frequency axis (a) and in time representation (b) for $T=40$ and $\rho=1$: results of MD simulation. Insert in (b) shows $Z(t)$ in double logarithmic scale. Graphs (c) and (d) show $|Z(\omega)|$ and $\ln|Z(\omega)|$ analytically continued into complex $\omega$-plain using multipoint Pade approximation built on top of $1400$ uniformly distributed knot-points in $\omega\in(0,200)$. The (half) width of the branch cut $\Delta \omega$ we mark by the blue bullet \textcolor{blue}{$\bullet$} on $Z(\omega)$. Characteristic time $\Delta t=2\pi/\Delta \omega$ we show by the red bullet \textcolor{red}{$\bullet$} on $Z(t)$ curve. }

\caption{ VAF at real frequency axis (a) and in time representation (b) for $T=200$ and $\rho=1$: results of MD simulation. Graphs (c) and (d) show 3D and density plots of $|Z(\omega)|$ in complex $\omega$-plain. Multipoint Pade approximation has been built on top of $1400$ uniformly distributed knot-points in $\omega\in(0,500)$. Again, the (half) width of the branch cut $\Delta \omega$ we mark by the blue bullet \textcolor{blue}{$\bullet$} on $Z(\omega)$. Characteristic time $\Delta t=2\pi/\Delta \omega$ we show by the red bullet \textcolor{red}{$\bullet$} on $Z(t)$ curve. }

166 — 1507.04576

\caption{Results of applying different methods on an egocentric photo-stream. Different bounding boxes show the tracking results of the \textcolor{green}{CT}, \textcolor{Blue}{LOT}, \textcolor{Cyan}{AMT}, \textcolor{yellow}{SPT}, \textcolor{magenta}{L1O} and \textcolor{red}{our} proposed approach. Occlusions can be observed in frame \#9 (a) and frames\#4 and\#9 (b).}

\caption{Results of applying different methods on an egocentric photo-stream. Different bounding boxes show the tracking results of the \textcolor{green}{CT}, \textcolor{Blue}{LOT}, \textcolor{Cyan}{AMT}, \textcolor{yellow}{SPT}, \textcolor{magenta}{L1O} and \textcolor{red}{our} proposed approach. Occlusions can be observed in frame \#5 (a) and frame\#6 (c).}

\caption{Results of applying different methods on an egocentric photo-stream. Different bounding boxes show the tracking results of the \textcolor{green}{CT}, \textcolor{Blue}{LOT}, \textcolor{Cyan}{AMT}, \textcolor{yellow}{SPT}, \textcolor{magenta}{L1O} and \textcolor{red}{our} proposed approach.}

\caption{Results of applying different methods on an egocentric photo-stream. Different bounding boxes show the tracking results of the \textcolor{green}{CT}, \textcolor{Blue}{LOT}, \textcolor{Cyan}{AMT}, \textcolor{yellow}{SPT}, \textcolor{magenta}{L1O} and \textcolor{red}{our} proposed approach. Occlusions can be observed in frame \#3 (a) and frame\#10 (b).}

167 — 1507.04752

\caption{Number density as a function of redshift for several galaxy samples considered in this paper. The magenta line corresponds to LOWZ targets (``TAR\_LOWZ''). The purple line corresponds to Legacy SDSS-II LRGs (``Legacy''). The blue line corresponds to\redm cluster members \citep[``RM\_CLUS'',][]{Rykoff:2014}. The green line corresponds to galaxies with $z_{\rm red}$ and $\chi^2_{\rm red}<5$ (``ZRED''). The solid red line corresponds to CMASS targets (``TAR\_CMASS'') and the dashed red line corresponds to CMASS galaxies from the LSS catalog (``LSS\_CMASS''). Grey dashed lines indicate the number densities of stellar mass threshold samples as a function of redshift estimated from the{\sc s82-mgc}.}

\caption{Stellar mass functions of the LSS\_CMASS sample in six redshift bins compared to the total SMF. Completeness is estimated via the ratio$c=\phi_{\rm A}/\phi_{\rm tot}$ where $\phi_{\rm A}$ is the number density of CMASS in a given redshift bin and $\phi_{\rm tot}$ is the number density of the total stellar mass function.}

\caption{Stellar mass functions of the LSS\_LOWZ sample in five redshift bins compared to the total SMF. Completeness is estimated via the ratio$c=\phi_{\rm A}/\phi_{\rm tot}$, where $\phi_{\rm A}$ is the number density of LOWZ in a given redshift bin and $\phi_{\rm tot}$ is the number density of the total stellar mass function. In this redshift bin, our double Schecter fit slightly under-estimates the total SMF at the high-mass end at $\log_{10}(M_*/M_{\odot})\sim11.8.$}

168 — 1507.04926

\caption[Diffusion constants, slice]{ (Color online) Diffusivities along the $x$- and $y$-direction, left and right panel respectively, for $\gi = 0.1$. Additional data of a 900-particle systems is marked by open symbols (short term: {\color{red} $\circ$}; long term: {\color{blue} $\diamond$}). For low shear rates, no diffusion is observed, corresponding to a reversible state. At a critical point $9.5 < \Gamma_{2,\mathrm{c}} < 10.5$ diffusions start to grow, both parallel and perpendicular to the shear, corresponding to an irreversible state. The inset in the left panel shows a magnification of the critical region. The short term diffusivity is slightly smaller than the long term diffusivity. Moreover, diffusion parallel to the shear is strongly enhanced and shows a different behavior: It grows linearly with increasing shear whereas diffusion in the perpendicular direction saturates. Boundary effects due to the system size seem to be negligible up to $\gs \approx 30.0$ since data points for $100$ and $900$ particles coincide well. \label{fig:ch2fig1}}

169 — 1507.05006

\caption{Summary of model fits to equation 6.1. %~\ref{fitfunction}: error Entries are prior range, posterior mean and standard deviation. The proton parameters are constrained by the \gray emissivities, while the lepton parameters reflect mainly the prior from synchrotron and direct measurements. The parameters are highly correlated and degenerate, so the resulting spectrum derived from the full posterior (Fig 1) is preferred to the individual parameters. The CR density $n_{ref}$ is multiplied by $(c/4\pi)$ to give a flux in the usual units quoted in experiments. $p_{ref}$= $10^5$ MeV for protons, $2\times 10^4$ MeV for leptons. }

170 — 1507.05199

\caption{Height-time observations and model predictions for CMEs 1, 2, 3 and 4. The height-time data points are denoted by diamond symbols along with error bars for COR2 \& HI data. The red dash-dotted line denotes the predictions of the model when it is initiated from the first observed data point, while the blue solid line denotes the model predictions when it is initiated from$\widetilde{h}_{0}$. In the case of CME 2, the green dash-dotted line indicates the model solutions when initiated from the start with a constant $C_{\rm D}$ of 0.1 while the brown dash-dotted line represents the model predictions using a constant $C_{\rm D}$ of 5.0. \label{fig1}}

\caption{Height-time plot for data and derived heights from the drag-only model for CMEs 5, 6 ,7 and 8. The height-time data points are denoted by diamond symbols along with error bars for COR2 \& HI data. The dash-dotted lines denoted the predictions of the model when it is initiated from the first observed data point, while the solid line denotes the model predictions when it is initiated from$\widetilde{h}_{0}$.\label{fig2}}

\caption{First and Third columns represent the model predictions for $C_{\rm D}$ (blue line) and $a_{d}/100$ (black line, in units of ${\rm cm}\,{\rm s}^{-2}$) for all events. For CME 2, the black line denotes $a_{d}/10^{4}$. Each plot is labelled with the CME number it represents. The second and fourth columns show model predictions for the Reynolds number (green line, in units of $10^{6}$) and $\gamma$ (black line, in units of $10^{-9} \,{\rm km}^{-1}$). In all cases except CME 2, the model is initiated from the height $\widetilde{h}_{0}$ (Table~\ref{tbl2}),{\color{blue} beyond which the CME can be considered to be solar wind drag-dominated.} The model for CME 2 is initiated from the first observed data point ($h_{0}$ in Table \ref{tbl1}). CME 2 is largely solar wind drag-dominated from the first observed data point. The discontinuities in some of the curves arise from discontinuities in the data at the interface between the fields of view of the different telescopes (e.g., between COR1 and COR2, between COR2 and HI). \label{fig3}}

171 — 1507.05443

\caption{Phase diagram of the Y$_2$O$_3$-Fe-Fe$_2$O$_3$ System at 1100\,$^\circ\textrm{C}$ \cite{Kitayama2004} and 1200\,$^\circ\textrm{C}$ \cite{Inazumi1981}. The region where stoichiometric YFe$_2$O$_{4-\delta}$ exists, is marked in \textcolor{red}{red} in the 1200\,$^\circ\textrm{C}$ diagram. (color online)}

\caption{Powder diffractograms of YFe$_2$O$_{4-\delta}$ for the non-stoichiometric single crystal (\textcolor[HTML]{000080}{NS SC}) with the magnetization in the inset of Fig.~\ref{fig:mag}, non-stoichiometric powder $(\delta > 0.05)$ (\textcolor[HTML]{008040}{NS P}) and almost stoichiometric powder $(0.03< \delta <0.04)$ (\textcolor[HTML]{F03232}{S P}), with the two step transition as in Fig.~\ref{fig:mag} but a difference at low temperatures between FC and ZFC. In light blue the reflection positions for the R$\bar{3}$m structure reported in \cite{Matsumoto1992a} are given. \\ \textit{Inset}: Dependency of the R$\bar{3}$m a/b-lattice parameter on the synthesis gas mixture. The given $\delta$ is determined from comparison of the magnetization data with those from \cite{Inazumi1981}, the correlation to the gas ratio is nonlinear \cite{Jacob2012}. For the single crystal the gas ratio was interpolated from the lattice parameter and the powder data, since it was synthesized in a CO$_2$~/~CO mixture. (color online)}

\caption{Magnetization of a 52\,mg\textit{stoichiometric} YFe$_2$O$_{4-\delta}$ single crystal $(0.00 \approx \delta \ll 0.03)$ grown in a CO$_2$~/~CO-ratio of 2.9$\pm$0.1, measured during field cooling (\textcolor[HTML]{000080}{FC}), field warming (\textcolor[HTML]{F03232}{FW}) after cooling in a field and after cooling without a field (\textcolor[HTML]{008040}{ZFC}). The powder data measured after 0.397\,T FC for\textcolor[HTML]{8B008B}{$\delta=0.00$} and \textcolor[HTML]{FF8C00}{$\delta=0.03$} is scaled by a factor of 2 for clarity and taken from \cite{Inazumi1981}. \\ \textit{Inset}: Magnetization of a \textit{non-stoichiometric} YFe$_2$O$_{4-\delta}$ single crystal, which was grown in a CO$_2$~/~CO-ratio of 2.6$\pm$0.1. (color online)}

\caption{a) Crystal structures (oxygen ions are omitted) and unit cells for the R$\bar{3}$m room temperature structure and the distorted P$\bar{1}$ for ($\frac{1}{4} \frac{1}{4} \frac{3}{4}$) as propagation vector~\cite{isodistort2}. (created with \cite{Momma:db5098}); b) Laue-diffraction image of YFe$_2$O$_{4-\delta}$ single crystal at 300\,K. c) Scan through$(\frac{1}{2} \frac{1}{2} \textrm{\footnotesize{13.5}})$ along $[00\ell]$ at 120\,K and$(\textrm{\footnotesize {0}}\,\textrm{\footnotesize {0}}\,\textrm{\footnotesize {18}})$ at 10\,K. The$\ell$-position of the 120\,K peak was corrected, based on the 10\,K UB-matrix. (color online)}

\caption{Energy dependence of the $(\frac{1}{2} \frac{1}{2} \textrm{\footnotesize{10.5}})$ superstructure reflection at 120\,K over the Fe$K$-edge, with two resonant features visible in the $\sigma \rightarrow \pi'$ channel at 7115.5\,eV and 7125.5\,eV. (color online)}

\caption{Polarization analysis at 7125.5\,eV on$(\frac{1}{2} \frac{1}{2} \textrm{\footnotesize{10.5}})$\at 120\,K, and of the direct beam. Solid lines are for the direct beam sinusoidal fits. The solid red and green lines are the expected behavior for an ideal Thomson scatterer. (color online)}

172 — 1507.05726

\caption{\scriptsize \color{blue} Examples from HandNet test set detections. The colors represent fingertips that are correctly located and identified. The white boxes indicate false detections with the error threshold chosen to be 1cm. The top two rows are trained and tested on non-derotated data. The bottom two are trained and tested on derotated data and then rotated back to the non-derotated space. The detections are overlaid on the IR image from the camera which is not part of the classification process. a) Successful examples for all methods. b) Representative challenging examples for which derotation enables better performance. c) Failure cases where derotation fails to improve the results.}

\caption{\scriptsize \color{blue} The data capture setup. a) 2mm magnetic sensors. The larger rectangular sensors are not used. b) A fingertip sensor inside the inner seam. c) Virtual model used for planning a multi-sensor setup. We only use 5 sensors. d) The RealSense camera rigidly fixed to the TrakStar transmitter. e) The back of the wooden calibration board where the glass sensor housings are firmly pushed through. f) The front of the calibration board where the glass sensor housings are visible on the corners as seen in the inset.}

\caption{\scriptsize \color{blue} The available data annotations after calibration. a) Color image. Illustrates a full hand setup for this work. The color is not used. b) The RGB axes indicate the measured location and orientation of each fingertip and the back of the palm. c) IR image(not used) overlaid with the labels generated from the raycasting described in Section \ref{sec:database}. d) IR image overlaid with the generated heatmaps per fingertip and the global orientation of the hand represented as an oriented bounding box (not used). }

\caption{\scriptsize \color{blue} Understanding derotation: We represent the space of poses by a non-uniform 2D region with representative hand poses. Red and green represent the pose-space covered by training images and testing images respectively. Each \RNum{1},\RNum{2},\RNum{3},\RNum{4} indicates one of the 4 possible combinations of training and testing for a machine learning method where the database remains fixed in size. The larger region indicates greater pose variability while the smaller represents less. Intuitively, by training on a space with low variance and testing in this same space (type \RNum{4}) we expect to see an improvement over the opposite (type \RNum{1}). Section \ref{sec:evaluation} supports this intuition. }

\caption{\scriptsize \color{blue} This graph shows the predicted value of all 9 coefficients of the hand orientation matrix in red relative to the ground truth in yellow. For clarity we order each ground truth coefficient monotonically and apply this reordering to the predicted results. The mean squared error for all the coefficients on the HandNet test set before and after SVD is 0.0271 and 0.0234 respectively.}

\caption{\scriptsize \color{blue} Synthetic and real examples of DeROT. a) The depth projection of the virtual hand before applying DeROT can be seen on the left wall of the cube representing the camera plane. The axis marked $r_{orient}$ is projected onto the camera plane and used in DeROT to define the angle $\alpha$. The purple circle contains the resulting image of the hand after applying derotation by angle $\alpha$. b) The top row of images are un-derotated. The bottom row have been derotated by $\alpha$ obtained by DeROT. Note that the thumb is consistently on the right of the image.}

\caption{\scriptsize \color{blue}{Results of our experimental evaluation for all experiment types described in Section \ref{sec:evaluation}. The experiment types \RNum{1},\RNum{2},\RNum{3},\RNum{4} are highlighted in red, pink, blue and cyan respectively. A result in bold indicates that it outperforms the baseline (\RNum{1}) shown in red. For each row pair (derotated training data vs non-derotated training), the underlined result is the better of the two. Procrustes consistently reduces the quality of fingertip detection. Conversely, DeROT outperforms the baseline \emph{for every experiment}. For all but one experiment, this improved performance is significantly enhanced by training on derotated data instead of original data. See Section \ref{sec:discussion}. The results from the Oracle serve as an upper bound achievable by derotation. }}

\caption{\scriptsize \color{blue} These graphs show typical precision to recall and precision to error threshold for thumb detection (using RDT and CNN on the HandNet test set. Each line indicates an experiment which is labeled in the legend using the experiment types from Section \ref{sec:experiments} and the derotation types Procrustes(\emph{a}), DeROT(\emph{b}), Oracle(\emph{c}). The baseline is in red. Training on derotated data and then applying DeROT or Oracle is in cyan or green respectively. Training on non-derotated data and then applying DeROT or Oracle is in magenta or black respectively. The average precision (AP) and precision at 1cm error (P@1cm) are shown for each thumb experiment.}

173 — 1507.05924

\caption{$v-i$ diagram for TDMA-based binary power talk: 2 VSC units (obtained with PLECS\textregistered{} simulation of the system shown in Fig.~\ref{2VSCMG}, $v^0=399V,v^1=401V,r_d^0=r_d^1=2\Omega,v_a^{\texttt{n}}=v_b^{\texttt{n}}=400V,r_{d,a}^{\texttt{n}}=r_{d,b}^{\texttt{n}}=2\Omega$.)}

\caption{$v-i$ diagram for FD-based binary power talk: 2 VSC units (obtained with PLECS\textregistered{} simulation of the system shown on Fig.~\ref{2VSCMG}, $v^0=399V,v^1=401V,r_d^0=r_d^1=2\Omega,v_a^{\texttt{n}}=v_b^{\texttt{n}}=400V,r_{d,a}^{\texttt{n}}=r_{d,b}^{\texttt{n}}=2\Omega$.)}

\caption{$v-i$ diagram for FD-based binary power talk: 3 VSC units (obtained with PLECS\textregistered{} simulation of the system shown on Fig.~\ref{3VSCMG}, $v^0=399V,v^1=401V,r_d^0=r_d^1=2\Omega,v_a^{\texttt{n}}=v_b^{\texttt{n}}=v_c^{\texttt{n}}=400V,r_{d,a}^{\texttt{n}}=r_{d,b}^{\texttt{n}}=r_{d,c}^{\texttt{n}}=2\Omega$.)}

174 — 1507.06073

\caption[]{\label{fig:beam-learn} Beam search for learning with different beam widths: \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {beam=10}; \draw[red] (-0.2, 0) -- (0, 0); \draw[red] (0, 0) -- (0.2, 0); \filldraw[fill=red,draw=black,radius=0.06] (0, 0) circle; \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {beam=20}; \draw[black!50!green] (-0.2, 0) -- (0, 0); \draw[black!50!green] (0, 0) -- (0.2, 0); \filldraw[fill=black!50!green,draw=black] (-0.07, -0.04) -- (0.07, -0.04) -- (0, 0.08) -- cycle; \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {beam=30}; \draw[blue] (-0.2, 0) -- (0, 0); \draw[blue] (0, 0) -- (0.2, 0); \draw[blue] (-0.07, 0.07) -- (0.07, -0.07); \draw[blue] (-0.07, -0.07) -- (0.07, 0.07); \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {exact.}; \draw (-0.2, 0) -- (0, 0); \draw (0, 0) -- (0.2, 0); \end{tikzpicture} \emph{Top}: Learning on $H_1$. \emph{Bottom}: Learning on $H_1 \circ_\sigma L_2$. The dashed line is the learning curve of the second-level cascade $H_2 \circ_\sigma L_2$. \KLcomment{what are the oracle error rates with different k?} \HTcomment{I've never calculated.}}

175 — 1507.06267

\caption{Our model reproduces qualitatively the nonequilibrium behavior seen in the growth experiments of \cc{Kim2009}. Results shown here are from growth simulations with (red) and without (black) the kinetic constraint. The fraction of colored components in notional solution that are blue is 0.5 (top) and 0.9 (bottom). We used the red-blue interaction energies obtained from \Tab{Crocker1}, with \#1 used in panel (a) and\#2 used in panels (b,c). Two aspects of the experiments of Kim{\it et al.}\c{Kim2009} are reproduced by our simulations in the presence of the kinetic constraint. 1) Kim {\it et al.} reported the absence of red impurities in structures grown slowly using the energy hierarchy \#1; the same is true in our simulations for a range of low growth rates (in panel (a), dashed lines are the experimental value from\c{Kim2009}). 2) Kim {\it et al.} reported the presence of a plateau in the `purity fraction' $\fb$ as a function of growth rate, between near-equilibrium and far-from-equilibrium growth regimes using energy hierarchy \#2. A plateau is also seen in our simulations (panel b; we do not expect the numerical value of this plateau to match that seen in\cc{Kim2009}, because our lattice model has a geometry different to that of the fcc crystals grown in experiment). In panel (c) we show that structures allowed to evolve post-growth evolve slowly towards equilibrium. \label{Fig:Crocker2}}

\caption{Growth (squares) versus growth+maturation (circles). Each graph displays the relationship between growth rates and outcomes from growth (squares) and growth followed by maturation (circles) for various solution concentrations ($\fb^\textrm{s}\in\{$\textcolor{red}{.2},.5,\textcolor{blue}{.8}$\}$; from top to bottom) and energy hierarchies--$\epsilon_\textrm{br}<0\equiv \epsilon_\textrm{rr}\equiv\epsilon_\textrm{bb}$ (a) and $\epsilon_\textrm{br}<\epsilon_\textrm{rr}\equiv0\ll\epsilon_\textrm{bb}$ (b). The lines in the graphs in (b) represent the magic number ratios obtained upon maturation. To the right of each graph is one snapshot of growth and one snapshot of growth+maturation at $\mu=4$. \label{FigGrowthVsMaturationAll}}

176 — 1507.07242

\caption{Examples of face search in first fold of the IJB-A closed-set 1:N search protocol, using ``templates." The first column contains the probe templates, and the following $5$ columns contain the corresponding top-$5$ ranked gallery templates, where \textcolor{red}{red} text highlights the correct mated gallery template. There are $112$ gallery templates in total; only a subset (four) of the gallery images for each template are shown.}

\caption{Top $10$ search results for the two Boston marathon bombers on the $80$M face gallery. The first two probe faces are of the older brother (Dzhokhar Tsarnaev) and the last three probe faces are of the younger brother (Tamerlan Tsarnaev). For each probe face, the retrieved image with \textcolor{green}{green} border is the correctly retrieved image. Images with the \textcolor{red}{red} border are near-duplicate images present in the gallery. Note that we were not aware of the existence of these near-duplicate images in the gallery before the search.}

177 — 1507.07685

\caption{(color online) Order parameters $R^{(1)}$ (solid black curve) and $R^{(2)}$ (dashed red curve) for the two coupled populations versus time. (a-c) Experimental measurements; (d-f) numerical simulations of the model (\ref{eq1}) with $m=10$. Initial conditions are BSCs for (a,b,d,e) and UCs for (c,f) and $N=15$. \red{The experiments in panels (a-b) (panel (c)) are carried out with $f = 160$~beats per minute (bpm) and $l= 17$cm ($f = 184$~bpm and $l= 25$ cm) and experimental time is measured in seconds.} }

\caption{\label{fig:swingsystem} \red{ (color online) Experimental setup: A sketch of the the mechanical system studied in~\cite{MartensThutupalli2013}, it is composed of two swings (A,B) coupled with a spring mechanism, each of which is loaded with $N=15$ metronomes.} }

\caption{(color online) (a) Average maximal LE \red{$\langle {\lambda}_1 \rangle$} vs $m$ for $N=100$: black circles (red squares) refer to BSCs (UCs). (b) Order parameter $R$ (black solid line) and finite time LE $\Lambda$ (red dashed line) for the chaotic population vs time $t$ for $N=200$ and $m=10$. The corresponding $P(\Lambda)$ (solid red line) is shown in (c) with the Gaussian fit (black dashed line). Inset: probability $p_0$ versus $N$. The values \red{$\langle {\lambda}_1 \rangle$} are obtained by following each realization for a time span $t=50,000$ and by averaging over 100 different initial conditions. }

\caption{(color online) (a) $\Lambda^{(*)}$ (symbols) versus $1/\ln(N)$ for $ 50 \le N \le 1600$ and the corresponding fit, the dot-dashed black curve with the shaded area denotes $\lambda_M(\infty)$ with its error bar. (b) Lyapunov spectrum for $N=100$, \red{the dashed red line is $\lambda^{(s)}$.} (c) Positive part of the Lyapunov spectra for various sizes, the dashed (green) line is $\lambda^{(c)}$. (d) Average life-times \red{$\langle \tau \rangle$} of the ICC vs $N$ for inertia $m=8$ (red squares) and $m=10$ (black triangles). Inset: \red{$\langle \tau \rangle$} vs $m$ for $N=30$ for ICCs. The blue dashed line in the main panel (inset) refers to a power law with exponent 1.60 (2.50). % Fitting per N>= 10 da valori degli esponent 1.60(5) (1.58 - 1.65) The values of \red{$\langle \tau \rangle$} are averaged over 200-3000 different realizations of BSCs and $m=10$ in (a-c). }

\caption{(color online) (a) Order parameters $R^{(1)}$ (black solid line) and $R^{(2)}$ (red dashed line) vs time; (b) imaginary vs real part of the complex fields $\rho^{(1)}$ and $\rho^{(2)}$. (c) Lyapunov spectra obtained for 3 different realizations of UCs. (d) Averaged components $\bar \xi_i$ (red diamonds) together with the corresponding averaged frequencies of the oscillators ${\bar \omega}_i \equiv \frac{\overline{ d \theta_i}}{dt}$ (black circles) for both populations. All the data refer to C2P states and to $m=9$, the system sizes are $N=200$ in (c) and $N=50$ in (a),(b),(d). \red{The time averages have been performed a time $T \simeq 5 \times 10^4$.} }

178 — 1507.07888

\caption{\highlight Illustration of pricing game with two SPs, homogeneous customers and unlicensed spectrum. The quantities $x_1, x_2$ and $X^w$ are the amount of customers served by SP1, SP2 and unlicensed spectrum, respectively. The areas of the rectangles $\pi_1, \pi_2$ are the revenues of SP1 and SP2, respectively. The area of the region between the delivered price and the inverse demand curve $P(q)$ is the consumer surplus, and the dashed area is total social welfare.}

179 — 1507.07934

\caption{%\footnotesize %\scriptsize {\color{red}\bf Top-left panel} show the average Compton-y map of 10,000 individual maps centered on the quasar host sampled from CS model at $z=1.4$. {\color{red}\bf The other five panels} show five randomly selected individual maps for five quasar halos. The pixel size is 0.034 arcmin.}

\caption{%\footnotesize %\scriptsize {\color{red}\bf Left panel} shows the mean Compton-y profile of quasars, including projection effects, for the CS (blue shaded region) and HOD (purple shaded region) model, respectively, for a synthetic sample of quasars with redshift distribution [$z=(0.1,3.0)$ with median redshift of 1.5] mimicing that of the observed sample used in \citet[][]{2015Ruan}. The radial profile of the observed quasars \citep[][]{2015Ruan} is overplotted as the green shaded region without dust correction. We have normalized the observed radial profile with that of the models at 30 arcmin radius, which corresponds to the ``background". {\color{red}\bf Middle panel} shows the mean Compton-y profile of a quasar, including projection effects, at three separate redshifts, $z=0.5$ (blue curves), $z=1.4$ (green curves) and $z=3.2$ (red curves), for the CS (solid curves) and HOD (dashed curves) model separately. {\color{red}\bf Right panel} is similar to the middle panel, except that only the quasar-hosting halo contributes to the thermal energy in the map, without considering the contribution from other halos due to projection effects. Also shown in black is the mean Compton-y profile in HOD model, without projection effects, with appropriate weightings in accordance with that of the observed sample used in \citet[][]{2015Ruan}. }

\caption{%\footnotesize %\scriptsize {\color{red}\bf Left panel} shows the stacked tSZ map of quasars with a median redshift of $1.5$ smoothed with FWHM=$1$~arcmin in the CS model. {\color{red}\bf Right panel} shows the same for HOD model.}

180 — 1507.08905

\caption{PSNR (SSIM) comparison on three test data sets among different methods. \textcolor{red}{Red} indicates the best and \textcolor{blue}{blue} indicates the second best performance. The performance gain of our best model over all the others' best is shown in the last row.}

181 — 1508.00409

\caption{\textbf{Left:} Function $u$ with cartoon-like derivatives. \textbf{Middle:} Cartoon-like first derivative of $u$ in vertical direction. \textbf{Right:} Plot on a logarithmic scale of the predicted decay of $O(n^{-2})$ for $n\to \infty$ due to \eqref{eq:decofCoeffs} \color{red}(dashed red line) \color{black} and actual decay of $(c(u)^*_n)^{\frac{1}{2}}$ \color{blue} (solid blue line)\color{black}. For better \color{black} }

182 — 1508.00948

\caption{The ECO catalog region in sky coordinates, with {\color{red}{the ECO mass-limited sample represented by points color-coded}} according to group halo masses estimated as described in \S \ref{envmet}. The RESOLVE-A region is outlined in red, and the region of overlap with the ALFALFA $\alpha.40$ catalog \citep{A40cat} is indicated by the purple crosshatched strips.}

\caption{The ECO catalog region in RA vs.\line-of-sight distance coordinates, in slices of$\sim$5 degrees in Dec, increasing from the bottom left panel. Galaxies included in the final ECO {\color{red}{mass-limited}} sample analyzed here are indicated by dots color-coded according to their group halo masses as in Fig.\\ref{skycov}, while galaxies outside this sample but present in our merged parent redshift catalog are indicated by grey dots. The outer limits of the ECO catalog ``buffer'' region are outlined in red, and the region of ECO we consider interior to this buffer is indicated in green. {\color{red}{Black lines in the upper left and bottom right panels indicate the approximate line-of-sight extent of two groups that extend significantly outside the ECO region and are subject to boundary completeness correction factors as described in \S \ref{ccs}. }} }

\caption{Illustration of the selection of the ECO catalog. Panel \emph{a} shows the $M_{r}$ and $M_{bary}$ distributions of the initial ECO catalog (dots and purple inset $M_{bary}$ histogram), where the horizontal red line indicates the $M_{r} < -17.33$ redshift completeness limit and the vertical green lines indicate the final mass cut we adopt to create an approximately baryonic mass limited sample to $M_{bary} > 10^{9.3} M_{\odot}$. {\color{red}{Panel \emph{b} shows the magnitude and cz limits of ECO, where light grey dots indicate the $M_{r} < -17.33$ sample and black dots indicate our final approximately mass-limited sample with $M_{bary} > 10^{9.3} M_{\odot}$ and group membership within the ECO volume.}} }

\caption{Illustration of the additional completeness of the {\color{red}{final $M_{bary}$-limited}} ECO catalog over the SDSS redshift catalog. The $M_{r}$ distribution of the ECO sample in our reprocessed magnitude system (see \S \ref{photdesc} for details) is indicated by the purple solid histogram, and the distribution of ECO galaxies that have SDSS redshifts is indicated by the black hashed histogram.}

\caption{The distribution of galaxies by {\color{red}{group halo mass in the mass-limited ECO, ECO+G, and RESOLVE-B samples, with completeness correction factors applied as described in \S \ref{ccs}.}} The red dashed vertical line indicates the group halo mass completeness limit for ECO, and the RESOLVE-B frequency histogram has been rescaled to match ECO at this limit. We have selected the ECO+G sample to have a similar environment distribution to the full ECO sample.}

\caption{Color vs.\stellar mass for the ECO catalog sample, where$(u-r)^{e}$ represents an internal extinction corrected color derived from the SED fitting code (see \S \ref{massmod}). {\color{red}{Dark grey points indicate individual ECO galaxies in our mass-limited sample, light pink points indicate ECO galaxies that do not enter this final sample, and density contours of the mass-limited sample distribution are shown in purple.}} The green line indicates our chosen red/blue sequence divider (\S \ref{rbseq}).}

\caption{Multiplicative completeness correction factors derived for each individual ECO {\color{red}{mass-limited sample}} galaxy (points) as a function of $r$-band absolute magnitude, model $g-i$ color, and {\color{red}{$r$-band surface brightness $\mu_{23.75r}$}}. Galaxies indicated by red points require further multiplicative correction factors on top of the factors shown, due to the loss of galaxies with extreme line-of-sight velocities outside the ECO definition. The majority of galaxies indicated by red points lie in the Coma Cluster, where the additional multiplicative correction factor reaches its maximum value of $\sim1.42$.}

\caption{Illustration of the traditional morphology-environment relation in the ECO and RESOLVE-B samples. Early and late type frequencies as a function of group halo mass in (a) ECO and (b) RESOLVE-B, with early and late type frequencies crossing over at $M_{halo} \sim 10^{13.5} M_{\odot}$ in the ECO sample. Greyscale dotted lines indicate the frequencies for central galaxies alone, which become poorly determined at high {\color{red}{group}} halo masses due to the small number of high halo mass groups (and thus centrals) present in the sample. In panel a, all frequencies are plotted at their ``expected'' value given the calibrated uncertainties in our semi-quantitative morphology classification method, described in \S \ref{morphclass}. The error bars in this panel are plotted as a combination of the calculated misclassification errors in each {\color{red}{group}} halo mass regime and the (binomial) counting statistics in each bin, while those in panel b reflect only the relevant counting statistics (assuming zero misclassification). Note that the early/late type frequencies for the two samples do not strictly agree in all bins, particularly near $M_{halo} \sim 10^{12} M_{\odot}$. Applying the quantitative classification approach to RESOLVE-B results in frequencies that more closely agree with the frequencies in ECO, although slight residual differences remain, perhaps resulting from group-to-group variations in typical properties.}

\caption{Illustration of morphology-environment relations in the ECO sample for separate high-mass ($M_{bary} > 10^{10} M_{\odot}$) and low-mass ($M_{bary} < 10^{10} M_{\odot}$) galaxy subsamples, indicated by thick and thin lines, respectively. Since few high mass galaxies inhabit {\color{red}{group}} halos below $\sim10^{12} M_{\odot}$, we do not plot the frequencies for high mass galaxies below this point. While the trends are qualitatively similar over most of the group halo mass range, the relations have different amplitudes for the low and high galaxy mass samples.}

\caption{{\color{red}{Group}} halo mass distribution for early and late type galaxies in the ECO sample, with corresponding grey dashed lines indicating the distribution for group central galaxies of each type. Panel \emph{a} shows all early and all late types together, and panel \emph{b} breaks down the early types further by red and blue sequence membership. Approximately 50\% of ECO early type galaxies occupy $M_{halo} < 10^{12} M_{\odot}$ environments, approximately 20\% of which are blue early types.}

\caption{{\color{red}{Group}} halo mass distribution for blue early-type and blue late-type galaxies in the ECO sample, with corresponding grey dashed lines indicating the distribution for group central galaxies of each type.}

\caption{Characteristic distribution of {\color{red}{group}} halo mass as a function of baryonic mass for different galaxy types. Solid lines indicate the running median group halo masses for centrals of each galaxy type, while dashed lines indicate the medians for satellites of each galaxy type. Error bars for each median point are estimated from the dispersion in properties in each bin. The background greyscale levels indicate the probability of inhabiting a particular {\color{red}{group}} halo mass at a given baryonic mass, as the histogram densities used to set the greyscale have been normalized to one in each baryonic mass bin (darkest points imply the highest probabilities).}

\caption{{\color{red}{Group}} halo mass distribution for blue early type galaxies in the ECO sample with different levels of HI gas. Corresponding greyscale dashed lines indicate the distribution for group central galaxies of each type. HI masses are derived as described in \S \ref{HIest}. Gas-dominated blue early types and blue early types with more moderate gas-to-stellar mass ratios primarily inhabit environments with group halo masses below $\sim 10^{12} M_{\odot}$, with $\sim$53\% of gas-dominated blue early types in $M_{halo} \lesssim 10^{11.5} M_{\odot}$ environments.}

183 — 1508.01065

\caption[Herringbone frequency time data and fits.]{(a) Frequency vs time data points for all detected bursts for both reverse (blue-green) and forward drifting bursts (orange-red). One line represents one burst, with 188 in total. (b) Linear fits to each burst detection. The green-red lines originate from $\sim$32\,MHz and are detections of the first set of bursts between 10:47:30--10:50:30\,UT. The blue-orange lines are the second set of bursts originating from$\sim$43\,MHz between 10:50:20--10:53:00\,UT. There are not equal numbers of positive and reverse drift bursts, indicating there is not a one-to-one correspondence of particle beams travelling toward and away from the sun, i.e. beams are bi-directional but not always in reverse and forward drift partners.}

184 — 1508.01483

\caption[Excitation functions]{(Color Online) Excitation functions for $\gamma p \to p \omega$ from CBELSA/TAPS ($\bullet$) for selected angle bins. For comparison, the CLAS data~\cite{Williams:2009yj} are represented by {\tiny \color{blue}$\square$}, the LEPS data \cite{LEPSeta} are represented by {\scriptsize \color{lepsaw}$\bigstar$} (for the $-0.8$ - $-0.7$ panel), and the Gie\ss en Lagrangian fit in \cite{Shklyar:2004ba} is represented by {\color{red}\rule[0.5ex]{0.1cm}{1pt} \hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt} \hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt}} . The Bonn-Gatchina PWA solutions are represented by green lines: {\color{green}\rule[0.5ex]{0.34cm}{1pt}} (full solution), {\color{green}\rule[0.5ex]{0.1cm}{1pt}\,\hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt}\,\hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt}} (solution without \threehalfplus{} partial wave). Statistical uncertainties are reported as vertical bars on each data point. The total systematic uncertainty for the CBELSA/TAPS data is shown as a black band at the bottom of each plot. Each plot is labeled with its range in cosine of the polar angle in the center-of-mass frame of the $\omega$ meson ($\cos \theta^{\,\omega}_{\rm c.m.}$). The horizontal axis is measured in the energy of the initial photon. An additional horizontal axis at the top of the figure shows the center-of-mass energy.}

\caption[Spin Density Matrix Elements Comparison]{(Color Online) Unpolarized spin density matrix elements in the Adair frame for selected energy bins. The $\bullet$ are the SDMEs extracted in this analysis with the total uncertainty for each data point represented as a vertical bar. The {\tiny \color{blue} $\square$} are the SDMEs published by the CLAS collaboration~\cite{Williams:2009yj}. The Bonn-Gatchina PWA solutions are represented by green lines: {\color{green}\rule[0.5ex]{0.34cm}{1pt}} (full solution), {\color{green}\rule[0.5ex]{0.1cm}{1pt}\,\hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt}\,\hspace{0.02mm}\rule[0.5ex]{0.1cm}{1pt}} (solution without \threehalfplus{}~partial wave). Each plot is labeled with its range in initial photon energy measured in the lab frame~(center-of-mass (W) energy).}

\caption[Spin Density Matrix Elements Comparison]{(Color Online) Polarized spin density matrix elements in the Adair frame for selected energy bins. The fitted Bonn-Gatchina PWA solution is represented by a {\color{green}\rule[0.5ex]{0.34cm}{1pt}} . The statistical and systematic uncertainties for each data point have been added in quadrature and are represented as a vertical bar. Each plot is labeled with its range in initial photon energy in the lab frame~(center-of-mass (W) energy).}

185 — 1508.02186

\caption{Static view of a 3D plot of handwritten digits projectd along first three MSIR directions, with points marked according to digit: \textcolor{blue}{$\circ$} = 0, \textcolor{green}{$\times$} = 6, \textcolor{red}{+} = 9.}

186 — 1508.02693

\caption{\coloronline The paths of $\ba'$ (red or dark) and $\bb'$ (blue or light) on the unit sphere for a loop of length 6.28 seen at time 500.0. Each vertex represents a segment with constant tangent vector. The tangent vectors have been projected from the unit sphere into the figure plane with the Mollweide projection (the same one usually used for cosmic microwave background maps). Kinks are represented by straight lines on the figure.}

\caption{\coloronline The paths of $\ba'$ (red or dark) and $\bb'$ (blue or light) on the unit sphere for a loop of length 8.29.}

\caption{\coloronline Close-up of the upper left of Fig.~\protect\ref{fig:join}.}

\caption{\coloronline 13197 consecutive segments of a long string at time 500.0, separated into 4 clumps by a hierarchical clustering algorithm. The clumps are shown with different colors (shading) and symbols. Dotted lines show the kinks that connect the clumps.}

\caption{\coloronline The leftmost clump of Fig.~\protect\ref{fig:long}, further broken down with threshold $40^\circ$}

\caption{\coloronline Magnified view of the bottom clump of Fig.~\protect\ref{fig:long}. Kinks that were present in the initial conditions are shown in blue (gray); those that were formed by later intersections are shown in black. The latter includes the kinks that go into and come out of the clump. The path of $\ba'$ sometimes jumps to a new value and then back again, leading to lines that appear to end.}

187 — 1508.03533

\caption{Evolution of restricted $\Lambda$-motis across the period 1995-2010. Panels a), b) and c) show the box plots of the $\Lambda$-motifs of sectors \textcolor{cyan}{$\bullet$} - manufacture of wood products, \textcolor{RoyalPurple}{$\bullet$} - basic metal industries and \textcolor{OliveGreen}{$\bullet$} - manufacture of food, beverages and tobacco across the considered time window. The mean of the year-specific BiCM-induced ensemble distribution is indicated as a yellow cross; light-blue dots represent the observed abundances of motifs. Panel d) shows the evolution of the motifs $z$-scores. Even at the macrosectors level, the peculiarity of the quinquennium 2003-2007 is apparent. Interestingly, the most robust early signals of the WTW structural change are detectable upon inspecting the most volatile economic sectors.}

\caption{Evolution of $z$-scores of restricted X-motifs across the period 1995-2010, measured on sectors \textcolor{cyan}{$\bullet$} - manufacture of wood products, \textcolor{RoyalPurple}{$\bullet$} - basic metal industries, \textcolor{OliveGreen}{$\bullet$} - manufacture of food, beverages and tobacco, \textcolor{Orchid}{$\bullet$} - manufacture of chemicals, \textcolor{Gray}{$\bullet$} - manufacture of fabricated metal products, for advanced economies (top panels) and emerging economies (bottom panels). Interestingly, the most robust early-warning signals of the WTW structural change are provided by emerging economies, for all the considered sectors.}

\caption{Evolution of restricted $\Lambda$-motis across the period 1995-2010. Panels a), b) and c) show the box plots of the $\Lambda$-motifs of sectors \textcolor{Violet}{$\bullet$} - manufacture of chemicals, \textcolor{MidnightBlue}{$\bullet$} - manufacture of fabricated metal products and \textcolor{OliveGreen}{$\bullet$} - textile across the considered time window. The mean of the year-specific BiCM-induced ensemble distribution is indicated as a yellow cross; light-blue dots represent the observed abundances of motifs. Panel d) shows the evolution of the motifs $z$-scores (lines joining the dots simply provide a visual aid). The macrosectors characterized by the highest values of internal similarity seem to be insensitive to the approaching crisis, showing remarkably stable (increasing or decreasing) tendencies throughout our whole temporal window.}

\caption{Evolution across the period 1995-2010 of the total nestedness - \textcolor{blue}{$\bullet$}, the row-specific nestedness - \textcolor{Magenta}{$\bullet$}, the column-specific nestedness - \textcolor{Plum}{$\bullet$} and the assortativity - \textcolor{BrickRed}{$\bullet$}. Panel a) shows the evolution of the observed abundances; panel b) shows the evolution of the corresponding $z$-scores, pointing out the peculiarity of the year 2003 in discriminating between a statistically significant phase (1995-2003) and a phase consistent with our null model (2003-2007). Lines joining the dots simply provide a visual aid}

188 — 1508.03663

\caption[]{The evolution, in the proof of Lemma~\ref{prop:40}, from $H$ to $H'''$ of a subgraph of $H$. Here $b_1, b_2, b_3 \in D$, $c_1, c_2, c_3 \in C$. Each \!\!\!%$*$ { \begin{tikzpicture} \draw[white] (0,0) -- (0,.001); \tikzstyle{A}=[draw,fill=black,star,star point ratio=3.8,star points=9,minimum size=3pt,inner sep=0pt] \draw (0,.10) node[A] () {}; \end{tikzpicture} } \!\!\! represents a vertex in$A$ adjacent to all three vertices incident to the face. (Black vertices have all incident edges drawn, but white vertices may have more incident edges.)}

189 — 1508.03928

\caption{Quantitative results using F-measure (higher is better) and MAE (lower is better). The best three results are highlighted in \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green}, respectively.}

190 — 1508.04016

\caption{Left: Order parameter $F = \frac{1}{N}\left|\sum_{i=1}^N {\boldsymbol \omega}_i\right|$ versus time showing a transition from the individual to the collective phase. The time period of a raid is indicated in blue beginning at $t_{dis}$, the time of food discovery, and ending at $t_{dep}$, the time of food depletion. Error bars represent one standard deviation. The letters a)-d) correspond to the snapshots of the raid in Figure~\ref{fig:trail} for a typical simulation. Right: Typical profile for the phermone concentration $c({\bf x},t)$ once the trail has been established and the collecitve state has been reached. The peak in $c$ occurs at the location of the food source and exponentially decays away from the trail. See Online Resource 2 for the evolution of the chemical concentration in time.}\label{fig:trans} \end{figure} % \begin{figure} % \centerline{\includegraphics[height=2.25in]{figs/chem_1fs}} % \caption{.}\label{fig:chem} %\end{figure} \subsection{Transition to a collective state} By introducing an order parameter \eqref{eqn:op} in Section~\ref{sec:op} that measures the coordinated behavior of each group we can study the transition to the collective state in time. Figure~\ref{fig:trans} shows a sharp transition to collective behavior after the time of food discovery $t > t_{dis}$. We notice that there is a time delay in the formation of a collective state of foragers due to the time elapsed from marking the food with pheromone and that pheromone diffusing out into the environment to be detected by others. This time delay is due to the interplay between the diffusion term, $\alpha \Delta c$, and the term governing the exponential decay, $\gamma c$ in \eqref{eqn:pher}. Figure~\ref{fig:trail}b) illustrates that locally near the food source where the pheromone has begun to diffuse the ants become attracted to the location of the food. As the pheromone diffuses out to the whole domain and the trail is laid more and more foragers become attracted. This can be seen by the steady increase of the order parameter $F$ for the foragers. While some may argue the returning ants reach a collective state in a trivial way due to the fact that the go directly to the nest, this does not account for the lane formation that will be discussed further in Section~\ref{sec:lane}. The raiding trail is considered to be formed when the order parameter for both the returners and the foragers is near one as illustrated in Figure~\ref{fig:trans}). Once the food is depleted at $t = t_{dep}$ we observe a rapid decrease of the order parameter. This is due to the fact that ants no longer lay chemical at that location and the pheromone evaporates exponentially fast. Since the chemical gradient has no bearing on the returning ants, the foragers deviate from collective behavior first. Those returning still must deliver the food they have to the nest along the home-bound vector before going on to other functions. After this time all the ants are foragers and in the absence of a chemical gradient or detection of a new food source all ants merely perform random walks returning to a disorder state. %In \cite{Cou03}, the authors take the average of the two groups and call it the flow, but here we keep them separate to investigate how the associated order parameter for each group evolves due to their different dynamics and interaction between groups. \begin{figure} %\centerline{\includegraphics[height=1.85in]{figs/histo_1fs_zones_final}} \centerline{\includegraphics[height=.9in]{fig4a}} \vspace{.075in} \centerline{\includegraphics[height=1.15in]{fig4b}\hspace{.05in}\includegraphics[height=1.15in]{fig4c}\hspace{.05in}\includegraphics[height=1.15in]{fig4d}} \caption{Formation of lanes along the pheromone trail for foragers and returners. Distances normalized by ant size $\ell=x_0$ (characteristic length). a) The trail is broken into three distinct segments with different distributions of ants. b) The average ant distribution over the entire trail shows bi-modal peaks occur at a distance between $.5\ell-1.5\ell$ from the center indicating two outside lanes of foragers (black) with returners (red) in the middle. The distribution of ants in each of the colored zones over time is found for c) foragers and d) returners. Error bars represent one standard deviation.}\label{fig:lanes} \end{figure} \subsection{Lane formation}\label{sec:lane} In addition to a transition to and from the collective state, we can also consider the local behavior along the trail. In particular, a histogram of the position of each ant with respect to its distance from the trail center is used to form an ant distribution function in the neighborhood of the trail. As can be seen in Figure~\ref{fig:lanes}, the foragers and returners naturally self-organize into lanes like cars on a highway or people in a crosswalk. Specifically, foragers who are driven by the chemical gradient occur between .5-1.5 ant lengths from either side of the trail with the highest probabilities forming a bi-modal distribution (see Figure~\ref{fig:lanes}b)). Whereas, returners who are driven by their desire to return to the nest as quickly as possible occur at the trail center with highest probability. We conclude, for the majority of the time along the trail, one lane of returners forms in the trail center and two lanes of foragers flank each side with equal probability. The formation of three lanes is consistent along the whole trail, but which class of ant is in the middle varies. By employing a microscopic model, unlike a continuum model, we can study different regions of the trail and focus on the local behavior (see Figure~\ref{fig:lanes}b)). Near the nest returning ants are in the center and foraging ants leave on either side with equal probability (Zone 1, red). In the central region (Zone 2, green) there is a crossover event where the populations switch lanes and foragers move toward the middle as they get closer to the food source. Here even 5 or 7 lanes of alternating classes of ants can be observed if the density of ants is large compared to the trail length. Near the food source (Zone 3, blue) the chemical gradient is strong and returners have equal probability of leaving with food on either side of the trail. Figures~\ref{fig:lanes}c) and d) focus on how the ant distribution changes between each zone. \begin{figure} \centerline{\includegraphics[height=1.5in]{fig5}} \caption{Removal of food over the course of time for a) one food source or b) two food sources at equal and unequal distances from the nest. Food depleted in 4-6 hours consistent with duration of raids from the experimental observations in \cite{Sch71,Sch40}.}\label{fig:food} \end{figure} How can ants form bi-directional traffic lanes? The model suggest it is the result of the excluded volume constraint and the break in symmetry between the social cues for the foragers and returners. In addition, the lane size introduces an effective length scale, which is dictated by the particle size manifested in the truncated repulsive potential $U$ defined in \eqref{eqn:truncated}. This global traffic behavior is consistent with the previous theory and/or experiment in \cite{Cou03,Dus09,Fra85}, where ants self-organize into lanes for optimal transport of food back to the nest. In addition, such traffic dynamics have also been recently observed in bacteria \cite{Ari13}. Unlike the model presented in \cite{Cou03}, we do not impose a directional preference for half the ants, which may artificially contribute to the formation of the bi-modal distribution in that work. Also, in \cite{Cou03} a turning parameter is used where outbound foragers have a higher avoidance rate, which essentially forces them to the outside of the trail. Instead, in this work, the excluded volume forces alone arising from first principles naturally sort the ants. Both models agree on the conclusion that lanes form due to the asymmetry in interactions between foragers and returners. The second natural question posed concerns the formation of three lanes as opposed to two. One explanation deduced from experimental observation in \cite{Cou03} is that a two lane flow would introduce a left-right asymmetry in the trail pattern not naturally present and thus limiting its efficiency. There may be another explanation. Our model suggests that the desire to return directly home with food outweighs the exponential decay of the chemical gradient away from the food source. In nature, when ant is encumbered with food it wants to return to the colony as quickly as possible (verified experimentally in \cite{MuhWeh88,Weh03}). When a foraging ant encounters a returning ant along a trail it is easier for the ant carrying nothing to move out of the way. This can also be seen in Figure~\ref{fig:lanes} by noting that the trail width is approximately the size of one ant. Alternatively, the first lane to appear and form is the central one for returners, which forms naturally in the middle to minimize the path back to the nest. Once foragers detect the pheromone the returning lane has already formed and they have no choice but to step away from the middle to avoid collisions until they get very close to the food source. When an ant returns to the nest with food and becomes a forager it has no bias to which side of the pheromone trail it will traverse. This leads to the bi-modal distribution of foraging ants along side the main trail. Similarly, when a returning ant leaves the food source it can be on either side of the trail explaining Figure~\ref{fig:lanes}d). \begin{figure} \centerline{\includegraphics[height=2.25in]{fig6}} \caption{Sample ant raiding simulations with foragers (purple) and returners (blue) where $N = 400$ and the food sources are equi-distant to the nest. Each arrow represents an individual ant's orientation ${\boldsymbol \omega}_i$. The black circles denote the nest, ${\bf x}_c = (20,0)$, and food sources, ${\bf x}_{f,1} = (80,20)$ and ${\bf x}_{f,2} = (80,-20)$. a) Initially ants are placed near the nest in non-overlapping positions with random orientation. b) The first food source is found and a trail develops similar to the case of only one food source. c) Once that food source is depleted random foraging commences once again until the other food source is found. d) A trail forms at the second food source. See Online Resource 3.}\label{fig:trail2eq} \end{figure} \subsection{Trail disappearance} Once the food source is exhausted the trail ceases to exist because the foraging ants no longer are attracted to it. This behavior is captured by imposing a count on the quantity of food items (e.g., 2000-3500). Once the food source is depleted no foragers can become returners and eventually the whole colony is composed of foragers looking for their next cache of resources. For some insight into how the food is efficiently broken down and returned to the nest see Figure~\ref{fig:food}a) for the quantity of food particles as a function of time. Since this function has essentially a constant decreasing slope after the trail forms, one could argue the system has reached the maximally efficient state and remains there until the food is gone. This provides further evidence for lane formation. If lanes did not form one would expect regions of little decrease in Figure~\ref{fig:food}a) representing congestion along the trail. After the trail disappears, $t > t_{dep}$, the simulations show that the ant distribution around the trail center for foragers becomes uniform and the lanes cease to exist (see Figure~\ref{fig:trail}d)). Also, after the food has been depleted and the disordered state commences, one may notice local areas of milling behavior similar to \cite{Ber06}. However, it is not well pronounced due to the presence of the random walk term in the dynamic equations. Now we wish to extend our study to make predictions about the behavior with multiple food sources. \begin{figure} \centerline{\includegraphics[height=2.25in]{fig7}} \caption{Sample ant raiding simulations with foragers (purple) and returners (blue) where $N = 400$ and the food sources are equi-distant to the nest. Each arrow represents an individual ant's orientation ${\boldsymbol \omega}_i$. The black circles denote the nest, ${\bf x}_c = (20,0)$, and food sources, ${\bf x}_{f,1} = (60,20)$ and ${\bf x}_{f,2} = (80,-10)$. a) Initially ants are placed near the nest in non-overlapping positions with random orientation. b) The first food source is found and a trail develops similar to the case of only one food source. c) Once that food source is depleted random foraging commences once again until the other food source is found. d) A trail forms at the second food source. See Online Resource 4.}\label{fig:trail2uneq} \end{figure} \subsection{Multiple food sources}\label{sec:two_food} In this section, the transition to the collective state and local lane formation in the presence of multiple food sources is investigated. Two main cases should be considered; namely, (i) two equidistant and (ii) two non-equidistant food sources. In principle different foragers can find each food source near the same time. Each will begin to deposit pheromone and return to the nest. Naturally, foragers begin to detect whichever pheromone is closer to their current location and follow the trail to that food source leading to the formation of two distinct trails. If the two food sources are at an equal distance from the nest one would expect the emergence of two near equivalent trails forming through the course of the raid. In contrast, if one food source is significantly closer, one would expect most foragers to detect that pheromone sooner and the vast majority would complete the raid on the first food source before moving to the second. Multiple foraging locations as well as the study of a trail network have also been considered in \cite{Amo14,Sum03}. Both cases can be understood by analyzing the PDE for the pheromone concentration \eqref{eqn:pher-new}. Since this equal is linear, multiple food sources can easily be considered by changing the righthand side to \begin{equation*} \sum_{j=1}^{M_1} qe^{\|{\bf x}_j(t)-{\bf x}_{f,1}\|^2}\delta({\bf x}-{\bf x}_j(t)) + \sum_{p=1}^{M_2} qe^{\|{\bf x}_p(t)-{\bf x}_{f,2}\|^2}\delta({\bf x}-{\bf x}_p(t)). \end{equation*} If one food source is visited more frequently, then more terms in \eqref{eqn:fund2} will direct ants toward that food site. We use simulations to study the distinct behavior among the two cases: (i) equidistant food sources (e.g., see Figure~\ref{fig:trail2eq} and Online Resource 3) and (ii) food sources at different distances (e.g., see Figure~\ref{fig:trail2uneq} and Online Resource 4). In the former case, the foraging ants are equally probable to find either food source while completing the random walk. We observe in simulations that both sites are visited initially, but the site that has more visitors eventually lures all the foraging ants due to the larger pheromone concentration. The typical three lane local dynamics of foragers and returners can be observed on each trail (see Figure~\ref{fig:lanes2}). In the case of food sources at unequal distances, the foraging ants find the closer food source first as they sweep across the computational domain. Once a food source is found in either case almost all foragers are attracted to this site, which exhibits behavior similar to the one food source case (see Figures~\ref{fig:trail2eq},\ref{fig:trail2uneq}). There is a period of random foraging again until the second food source is found. \begin{figure} \centerline{\includegraphics[height=1.65in]{fig8}} \caption{Order parameter $F = \frac{1}{N}\left|\sum_{i=1}^N {\boldsymbol \omega}_i\right|$ versus time showing a transition from the individual to the collective state. Left: Two equally spaced food sources. Letters correspond to Figure~\ref{fig:trail2eq}a)-d). Right: Two no-equidistant food sources. Letters correspond to Figure~\ref{fig:trail2uneq}a)-d). Error bars represent one standard deviation. Observe in the case of equally spaced food sources the raids last roughly the same amount of time indiciating a maximally efficient state has been reached and raid time only depends on the trail length.}

191 — 1508.04518

\caption[]{\small Example of a symplectic multi-billion year integration of the Solar System with \hnb\(run \NSY). Results shown are from symplectic integrator only (4-th order with corrector and Jacobi coordinates), not from \BS\integrator (see text). (a) Relative energy error,$|\D E/E| = |(E(t)-E_0)/E_0|$. After the two reductions in integrator timestep ($t_i$, $i = 1,2$), solid and dashed lines show $|\D E/E|$ relative to $E(t_i)$ and relative to absolute $E_0(t_0 = 0)$, respectively. (b) Relative angular momentum error, $|\D L/L| = |(L(t)-L_0)/L_0|$. (c) Change in Mercury's and Earth's semi-major axes, $\D a/a = (a(t)-a_0)/a_0$. (d) Mercury's and Earth's eccentricity (\eM, \eE). (e) Inclination. Note that $|\D E/E| < 10^{-10}$, even at $\eM \simeq 0.8$. When oscillations in $|\D E/E|$ and/or shifts in \aM\occurred ($\eM\gtrsim0.8$), integrations were continued with \ms's \BS\algorithm\tcx{(see Figs.~\ref{FigFail}, \ref{FigBSm6}, \ref{FigFailAll}).} }

\caption[]{\small Mercury's maximum eccentricity per 1~Myr bin from the 1,600 symplectic 5-Gyr integrations. Results shown are from \hnb's symplectic integrator only \citep{rauch02}. In ten solutions, \eM\crossed 0.8, after which the integration was continued with\ms's \BS\algorithm\citep{chambers99} (Fig.~\ref{FigBSm6}). Note that in R0840 (arrow), $\eM < 0.8$, no shifts in \aM\occurred, and$\max|\D E/E| \simeq 8\e{-11}$. }

\caption[]{\small 120 Myr of \BS\integration of ten solutions, initiated when oscillations in$|\D E/E|$ and/or shifts in \aM\occurred in the symplectic integrations ($\eM\gtrsim0.8$, see text). (a) Relative error in energy and (b) angular momentum. (c) Mercury's semi-major axis and (d) eccentricity. Note solutions R0043 and R1231. }

\caption[]{\small Maximum values per 1-Myr bin of Earth's slowly changing orbital elements. (On gyr-time scale, argument of periapsis, longitude of ascending node, and mean anomaly may be considered 'fast angles'.) For better visualization, only every 20th of the 1,600 solutions are plotted plus all runs with $\eM > 0.8$ for $t = 2.5-$5~Gyr. The 1,600 solutions used initial conditions that differed only by 1.75~mm in Mercury's initial radial distance between every two adjacent orbits. (a) Maximum semi-major axis, (b) eccentricity, and (c) inclination. All three elements remained within narrow bands in the 1,600 solutions: $\max(\aE) < 1.0005$~AU, $\max(\eE) < 0.15$, and $\max(\iE) < 5.1\deg$. The runs with elevated \aE\and\eE\are solutions with large increases in Mercury's eccentricity. For example,\aE\reached 1.00043~AU in\RXA\during a high-\eM\interval, which occurred$\sim$8~Myr before Mercury collided with Venus. After Mercury had been merged with Venus, the system behaved very stable. }

192 — 1508.05046

\caption{\em \small An example of $\gamma_\t(x)$ in the range of $[0,1]$, with five distinct values (\textcolor{red}{$\diamond$}) corresponding to $[0.1, 0.5, 0.7, 0.8, 0.9]$.}

\caption{\em \small Examples of function $\phi_{\t,\lambda}(c)$ for different $\lambda$ values and the same set of $\t=[0.1, 0.5, 0.7, 0.8, 0.85]$ (\textcolor{red}{$\diamond$}). Left: soft-rounding ($\lambda=0.6$). Right: rounding ($\lambda=1.1$).}

193 — 1508.05903

\caption{Examples of traces in the liquid (163\,K, blue trace) and the solid (157\,K, red trace) phase of xenon. The applied electric field in these examples is 500\,V/cm. The glitch peaks at 40\,$\pm 0.5 \mu$s are induced by the high voltage trigger of the pulsed light source unit. The anode signal in the solid xenon is faster than that in the liquid xenon. The amplitudes of the cathode signals are similar in liquid and solid xenon. The functional form fit results are overlaid on the traces in light-blue curves. The charge drift distance between the photocathode (trigger signal) to the start time of the anode signal is 8.0\,cm.~\label{fig:tracecomp}}

\caption{Electron drift velocity in liquid and solid xenon for various electric field strength. The measured electron drift velocities in liquid xenon before the solidification are in magenta squares ({\textcolor{magenta}{\tiny$\blacksquare$}}). The measurements in solid phase for the first trial of solidification are in red triangles ({\textcolor{red}{\tiny\ding{115}}}). The measurements in liquid phase after melting the xenon are in blue circles ({\textcolor{blue}{\tiny\ding{108}}}), where a slight offset for the electric fields is just for display purpose. The second trial of solid xenon measurements after refreezing the xenon are in magenta down-triangles ({\textcolor{magenta}{\tiny\ding{116}}}), where the electric fields are off-set just for display purpose.. The previous measurements by Miller {\it et al.} in thin-layers (145$\mu$m to 228$\mu$m) of liquid and solid xenon are shown in stars ({\ding{72}}) and blank squares ($\square$) respectively~\cite{Miller:1968zza}. The red line ($u_S$) indicates the sound velocity in solid xenon (0.110\,cm/$\mu$s) at 157\,K. The blue line ($u_L$) indicates the sound velocity in liquid xenon (0.065\,cm/$\mu$s) at 163\,K. The red dashed curve shows the effective Shockley theory. See text for details.\label{fig:evelocity}}

194 — 1508.05912

\caption{UV cure glues dispensed with a microlitre pipette on red plastic (left, DYMAX\textregistered~6-621) and a fully automated glue dispenser (right). While a pipette allows only the application of a single glue spot, an automated glue dispenser can reproduce the original five-dot glue pattern for DYMAX\textregistered~6-621 and LOCTITE\textregistered~3525.}

\caption{Backside of an ASIC glued with LOCTITE\textregistered~3504, removed after curing with UV light guides connected to a mercury arc lamp. The glue layer is broken around three corners of the ASIC and shows additional cracks toward the ASIC centre. No regions of uncured glue were found.}

\caption{List of materials used in components for silicon strip detector modules with coefficients of thermal expansion (CTE). The adhesives show significantly higher CTE than constituents of components. DYMAX\textregistered~UV cure glues with no specified CTE can be assumed to have values of CTE comparable to LOCTITE\textregistered~UV cure glue because of their similar viscosity, density and composition.}

\caption{Peak shear forces for ASICs glued with silver epoxy glue (TRA-DUCT\textregistered 2902) and possible replacement glues, after irradiation and thermal cycles. For each glue, two ASICs were used for shear tests on each hybrid (columns I and II). Numbers in parenthesis represent measurements where an ASIC was damaged instead of being removed. POLYTEC\textregistered~UV 2133 and 3M\textregistered~5590H were found to be insufficient.}

\caption{Overview of test results of seven glues under investigation: three glues were rejected as possible replacements. Among the remaining four UV cure glues DYMAX\textregistered~3013 and 6-621 are considered preferable because of a lower toxicity classification.}

\caption{Comparison of the temperature profiles measured on a hybrid glued with silver epoxy glue and a hybrid glued with UV cure glue (DYMAX\textregistered~6-621) for unirradiated (white background) and irradiated (grey hatched background) hybrid areas. Temperature differences between ASICs and hybrid were found to be consistently larger for the UV cure glued hybrid, compared to the silver epoxy glued hybrid, by {\unit[2-5]{$^{\circ}$C}}.}

195 — 1508.06163

\caption{Visual comparisons of road area extraction results with four comparing methods. There are two rows and six columns of subfigures. Each row shows results on one dataset. From left to right: (a) original image, (b) result of Huang \cite{journals_ijrs_Huangxin09}, (c) result of Miao \cite{journals_lgrs_MiaoSZW13}, (d) result of $\textrm{Shi}^{a}$ \cite{journals_lgrs_ShiMWZ14}, (e) result of $\textrm{Shi}^{b}$ \cite{journals_tgrs_ShiMD14}, (f) result of $\textrm{Proposed}^{3}$. From (b) to (f), {\color{green}{true positive (TP)}} is marked in green, {\color{red}{false positive (FP)}} in red, {\color{blue}{false negative (FN)}} in blue. (Best viewed in color)}

\caption{Quantitative comparisons with state-of-the-art methods in two datasets, where the {\color{red}{\textbf{red values}}} marked in bold are the best and the bold {\color{blue}{\textbf{blue values}}} are the second best. (Best viewed in color)}

\caption{Quantitative comparisons with state-of-the-art methods in Data $\#1$, where the {\color{red}{\textbf{red values}}} marked in bold are the best and the bold {\color{blue}{\textbf{blue values}}} are the second best. (Best viewed in color) \label{centerline_table1}}

\caption{Quantitative comparisons with state-of-the-art methods in Data $\#2$, where the {\color{red}{\textbf{red values}}} marked in bold are the best and the bold {\color{blue}{\textbf{blue values}}} are the second best. (Best viewed in color) \label{centerline_table2}}

196 — 1508.06308

\caption{Average number of electrons in the gap $<\rm EIG>$ (red circles), average number of electrons per bunch $<{\rm EPB}>$ (blue squares) \red{and averaged frequency $<{\rm f\ [THz]}>$ (black triangles). Markers are slightly shifted horizontally for clarity.} %better visualization.} }

197 — 1508.06716

\caption{(color online) (a)-(d)~Steady-state slip velocities $v_s$ measured at the rotor (\textcolor{red}{$\square$}) and at the stator (\textcolor{blue}{$\blacksquare$}) for different temperatures $T=17$, 20, 26 and 34$^{\circ}$C. Gray lines are the best power-law fits of the data, $v_s=B(\sigma-\sigma_c)^{p}$, where $\sigma_c$ is fixed to the yield stress inferred from the Herschel-Bulkley fit of the flow curve (Fig.~\ref{fig1}) and $B$ and $p$ are free parameters. (e)~Exponent $p$ vs temperature $T$ obtained from the previous fitting procedure ($\bullet$) and by using the value of $\sigma_c$ that minimizes the $\chi^2$ of linear fits of $\ln v_s$ vs $\ln(\sigma-\sigma_c)$ when $\sigma_c$ is varied ($\circ$). In the former case error bars are computed from the uncertainty on $\sigma_c$, whereas in the latter case error bars correspond to the standard deviation associated with the least-square method. The horizontal dashed lines emphasize $p=1$ and 2. \label{fig3}}

198 — 1508.06722

\caption{(a) Cross section of a spin-guide. Two circles show the position of the wires, where the inner circle and cross inside each circle signifies the direction of the current. The horizontal line is the plane of the spin sheet. $2w_g$ is the separation between wires and $d$ is the distance between the spin sheet and the wires. (b) Onsite energy of spins inside the sheet due to individual (\\square \and\\circ\) and combined (\\ccirc \) magnetic field of two current carrying wires. Two vertical (dotted) lines show the position of the wires. The two wires form a potential well with depth $\varepsilon_{\textrm{min}}$, which can guide magnons.(c) A pseudocolor plot of the depth, $\varepsilon_{\textrm{min}}$, due to the magnetic field of the wires, located at the center of the potential. The black region is the region where the potential starts to split into two separate potentials and therefore leads to more complicated dynamics. The inset shows the shape of such a split potential with the parameters that lie inside the black region($d = 80\textrm{a}, w_g = 400\textrm{a}$). The white line marks the position of the maximum potential depth.}

199 — 1508.06848

\caption{Phase coherence and intervalley scattering time. Phase coherence (\textcolor{red}{$\blacksquare$}) and intervalley ($\bullet$) scattering time obtained by fitting with equations of Ref.~\onlinecite{Kuntsevich2007}. The black circles ($\circ$) are obtained with Hikami formula for $B<0.15$ T. The dotted red line represents the $T^{-1}$ dependence and the blue solid line represents the prediction of ZNA theory with $\mathcal{F}_0^\sigma$=-0.67. The insert shows the experimental magneto-conductivity (black curves) and result of the fit (red curves) for different temperatures (T=0.3 K, 1.5 K, 3 K, 4 K, 5 K, 6 K from top to bottom.). The arrow indicates the crossover field $B_v\approx0.25$ T (See the text). \label{WL}}

200 — 1508.06944

\caption{\textbf{Dynamics of finite $N$ Networks}. \textbf{A} Mean activities of the four populations after initialization at a specific state along the approximate attractor. Blue (red) traces are used for the excitatory (inhibitory) populations. Dashed and solid lines are used to distinguish between the two subnetworks. \textbf{B} Projection along the approximate attractor, shown for the same simulation as in A. \textcolor{myblue}{\textbf{C} Projected dynamics for two initial conditions: $X(0)=0.05$ (blue shaded area) and $X(0)=0.08$ (grey shaded area). Black lines are averages over 50 trials. Shaded areas represent the standard deviation measured over 50 independent trials.} \textbf{D} Projection of the mean activities on the $m_1$-$m_3$ plane in resting state activity. \textbf{E} Dynamics of the projection $X$ along the special direction (black), and a projection on a perpendicular direction (red) in the same simulation as in D. In A,B and C $N=1.5\times 10^5$, $K=500$ and $\tilde{J}\simeq 1.77 $. In D and E $N= 10^5$, $K=1000$ and $\tilde{J}\simeq 1.69 $. Other parameters are as in Fig.~\ref{fig:finite_K}. }

201 — 1508.07822

\caption{ \label{switch} The \textcolor{red}{input-output relation} follows the lower stable equilibria until it becomes unstable through a fold (also called saddle-node) bifurcation, then jumps to the upper stable branch. Hence such an input-output relation is a good switch. }

202 — 1508.07926

\caption{The absolute difference of points between the shaded and not shaded regions is at most \tred{$(n-2)/3$ (left) or $(n-3)/3$ (right).}}

203 — 1509.00052

\caption{ \label{fig:double scaling}(Color online) Scaling behavior of two \bubbles\arranged perpendicularly to an incident pulse for various radii and initial separations. The (\redTriangle, \blueCircle) symbols on each axis denote data associated with that axis. The \redTriangle\follow a regression of$\Delta \abs{\vb{v}}_\text{d} = \num{0.250754} d_{12}^{-3.00077}$, and the \blueCircle\follow$\Delta\abs{\vb{v}_\text{max}}_\text{r} = \num{3.13328e-5} a_0^{2.99814}$. These trends strongly indicate dominant dipolar interactions between \bubbles. }

\caption{ \label{fig:isosurface}(Color online) Isosurfaces of velocity potential (arb.~units) calculated by evaluating the $\hat{\mathcal{S}}$ and $\hat{\mathcal{D}}$ terms in \cref{eq:reduced Kirchoff-Helmholtz} for a $N = 16$ particle simulation. Red, blue, and yellow surfaces denote regions of positive, negative, and zero potential, with holes appearing due to intersections with the bounding box. The inset box shows the three dimensional arrangement of the \bubbles\superimposed with their velocity vectors, as well as several positive and negative potential isosurfaces. Rendered with VisIt\cite{VisIt}. }

204 — 1509.00208

\caption{\red{Lattice potential structure for a non-vanishing external electric field.}}

205 — 1509.00298

\caption{The effect of the heavy resonance mass-shell constraint is demonstrated using the mass variables $M_{T2}$ and $M_{2Cons}$ considering one antler event in the invisible momentum component space. The region represented by the color gradient is the maximum among two transverse masses coming from two decay chains. Corresponding contours are shown with dashed lines. Following Eq.~\protect \ref{mt2}, the minimization of this quantity, the $M_{T2}$, is represented by the filled circle \tikzcircle[black,fill=gray]{3.5pt}. In the same plot, the solid lines (of same colors) are delineating the corresponding contours for the $(1+3)$-dimensional new variable with heavy resonance mass-shell constraint as in Eq.~\protect \ref{m2Cons}. Note that only the contour lines are shown in this case, not the color gradient as in the transverse mass case. The minimum of these solid contours is represented by the $M_{2Cons}$, displayed by circle plus $\oplus$ in the same figure. The $G$ mass-shell constraint restricts the invisible momenta, making the region shrink, as depicted by the dashed and solid lines. The white dashed (solid) line represents the equality of transverse-mass (mass) of parents and this equality line is also moving towards higher values because of the constraint. The red star $\bigstar$ is the position of true transverse momenta of invisible particle. The mass spectrum we choose is ($m_G$, $m_{P}$, $m_{\chi}$) = (1000.0, 200.0, 100.0) in GeV and the trial invisible particle mass $\tilde{m}_\chi$ we took for this plot is 10.0 GeV}

206 — 1509.00750

\caption{ Relative efficiency of the $L_p$-norm estimates (Y-axis) for different noise types. The efficiency is calculated as ratio of the $L_2$-estimate variance for the EST-design to other $L_p$-estimates for all designs.\\ \full EST, \dotted ESD$_{_\textrm{\tiny{RF}}}$, \longbroken ESD$_{_\textrm{\tiny{FF}}}$. }

207 — 1509.00849

\caption{Left and Right Bundle Brunch Block --~Healthy case in \textcolor{myred}{red}, LBBB in \textcolor{myblue}{blue} and RBBB in \textcolor{mygreen}{green} (voltages (mV) versus time (ms))}

\caption{Bachmann's Bundle Block --~Healthy case in \textcolor{myred}{red}, BBB in \textcolor{myblue}{blue} (voltages (mV) versus time (ms))}

\caption{Wolff-Parkinson-White syndrome --~Healthy case in \textcolor{myred}{red} and WPW in \textcolor{myblue}{blue} (voltages (mV) versus time (ms))}

\caption{ECGs obtained with different ionic models --~Courtemanche/MV ionic model in \textcolor{myred}{red} and Mitchell and Schaeffer model in \textcolor{myblue}{blue} (voltages (mV) versus time (ms))}

208 — 1509.01276

\caption{Comparisons of various total magnitude estimates from the UKIDSS LAS catalog to MAG\_AUTO remeasured with SExtractor. UKIDSS Petrosian magnitudes (Petro), Kron magnitudes (Kron), isophotal magnitudes (Iso), Hall magnitudes (Hall), and corrected ``total'' AperMag6 magnitudes (Aper6) are displayed. Hatched histograms refer to offsets for isolated sources while solid, filled histographs show results for deblended sources.\label{fig:maghist_remeasure}}

\caption{Comparisons of three photometric redshift estimators to available spectroscopic redshifts in Stripe 82. The comparison is limited to $i < 22.5$ and $0.01 < z_{\rm spec} < 0.8$. The left and middle panels are from the \redmapper{} project (Rozo et al., in preparation), while the right panel compares neural-network \photozs{} from \citet{reis12}. The 3$\sigma$-clipped dispersion is listed in each panel along with the fraction of catastrophic outliers defined by $\abs{\Delta z} > 0.1$. Contours are plotted at high data densities with 0.3 dex logarithmic spacing in the left and middle panel and 0.4 dex in the right panel. The 1-to-1 relation is plotted in each panel as a thin light grey line. \label{fig:photoz_specz}}

\caption{Fractional contribution of different redshift results as a function of $M_*$ and redshift to the ``best'' estimator, \zbest{}, appropriate for bright galaxies ($i \lesssim 22.5$) in the \cat{}. The red line ($z_{\rm spec}$) indicates \specz{} from SDSS. The labels $z_{\rm RM}$ and $z_{\rm red}$ refer to \redmapper{}. ANNz \photozs{} from \citet{reis12} are labeled $z_{\rm phot}$ and shown in green. See text for details. This figure is reproduced and described further in Paper II. \label{fig:zbest_fraction}}

209 — 1509.01287

\caption{\label{fig:scale} \cred{Evolution of classification performance with number of scales. Results obtained from $30$ Monte Carlo runs with different training sets of $10$ randomly selected samples per class of the image in Fig. \ref{fig:natural}. The variation of performance for more than $4$ scales is negligible. }}

\caption{ \label{fig:LORSALlambda} LORSAL parameter analysis. Effect of $\lambda$ on the overall accuracy values and sparsity of $\W$. \cred{Mean accuracy (in black), standard deviation (in gray), overlapped with the results for all images.} Note the three zones of accuracy behavior: no effect, increase, decrease. The maximum overall accuracy ($66.4 \%$) is obtained for $\lambda = 10$ with a value of relative regressor sparsity of $0.352$.}

\caption{ \label{fig:quality} Variation of quality of classification $Q$ with the contextual index $\alpha$ and the rejection threshold $\rho$ for four different training sets. Adjacent contour lines correspond to a $0.01$ variation of $Q$. \cred{It is clear a shift to lower dependency on rejection and contextual information as the size of the training set, and consequently the classifier performance, increases.}}

\caption{ \label{fig:class_res} Example of classification results for H\&E stained samples of teratoma imaged at$40X$ containing multiple tissues: Image 1 (first row) background (light pink), smooth muscle (dark pink), gastro intestinal (purple), mature neuroglial (light brown), fat (dark brown); mesenchyme (light blue); Image 2 (second row) background (light pink), fat (dark brown), mesenchyme (light blue), skin (green); Image 3 (third row) mesenchyme (light green); bone (dark blue). Rejected partitions are shown in black. The training set consists of: \cred{$5$ randomly chosen partitions per class (roughly $0.6\%$ of total) for image 1, } $120$ randomly chosen partitions (roughly $3\%$ of total) \cred{for image 2}, \cred{$20$ randomly chosen partitions (roughly $ 0.5\%$ of total) for image 3}; with the $\lambda$ parameter set to $5$. %Nonrejected accuracy: Image 1/2/3 $87\%$/$90\%$/$96\%$. Classification quality: Image 1/2/3 $84\%$/$89\%$/$96\%$. Rejection quality: Image 1/2/3 $11.6$/$25.4$/$27.9$. Rejected fraction (partitions): Image 1/2/3 $11\%$/ $5\%$/$2\%$. }

210 — 1509.02348

\caption{The hyperplane $H$ (plain line) produces the same classification (into \textcolor{blue}{$\g +$} and \textcolor{red}{$\g\circ$}) as the hyperplane $H_S$ (dashed line) obtained by a translation (dotted line) and a rotation of $H$ such that it passes through exactly 2 points of $S$ %$\g x_1$ and $\g x_2$ (\textcolor{magenta}{$\g *$}). \label{fig:prophyperplane}}

211 — 1509.03580

\caption{\small Dynamo view of trajectories of the Lorenz system. The exact system is shown in black ($-$) and the sparse identified system is shown in the dashed red arrow ({\color{red}$--$}).}

212 — 1509.04567

\caption{$\color{blue} \times$: $\nu = 10^{-1}$; $\square$: $\nu = 10^{-2}$; {\color{blue} dotted line}: $\nu=10^{-1}$ (ideal); dashed line: $\nu=10^{-2}$ (ideal).}

213 — 1509.04930

\caption{(Color online) Field-driven transition from HQV to FV due to confinement in a square mesoscopic sample of size 8$\xi\times$8$\xi$. (a) Energy of the state \emph{f} of Fig. \ref{free_dk00} (a), along with some of its neighboring states. (b) Second order derivative of the energy with respect to the external field showing three distinct states indicated by circular, squared and triangular symbols. The corresponding components of the superconducting order parameter are shown in panels (\textcolor{blue}{$\bigcirc$}), (\textcolor{red}{$\square$}) and (\textcolor{green}{$\triangle$}). Displayed quantities are logarithmic contour plots of $|\psi_x|^2$ and $|\psi_y|^2$ in left and central columns, respectively, while the cosine of the phase difference is shown at the right column.}

\caption{(Color online) Temporal vortex-skyrmion transition in a square mesoscopic sample of size 12$\xi\times$12$\xi$. Panel (a) shows the free energy of states \emph{i} and \emph{j} containing 10 and 12 fractional vortices per component, respectively. The energy of state \emph{i} is discontinuous at $H\approx 1.06H_{c2}$ reflecting a first order transition. Panel (b) shows the temporal evolution of the energy at the latter transition. Three states, initial, intermediate and final are denoted by circle, square and triangle markers, respectively. The components of the superconducting order parameter corresponding to each state are shown in panels (\textcolor{blue}{$\bigcirc$}), (\textcolor{red}{$\square$}) and (\textcolor{green}{$\triangle$}).}

\caption{(Color online) (a) Free energy as a function of the anisotropy parameter $\delta_k$ with the external magnetic field fixed at $H=0.530[H_{c2}]$. Three distinct states are indicated by circular, square and triangular markers (\textcolor{blue}{$\bigcirc$}, \textcolor{red}{$\square$} and \textcolor{green}{$\triangle$}). (b) and (c) show line profiles of $|\psi_+|^2$ and $|\psi_-|^2$, respectively, along the diagonal line ($y=x$) corresponding to the states of panel (a). Columns (\textcolor{blue}{$\bigcirc$}), (\textcolor{red}{$\square$}) and (\textcolor{green}{$\triangle$}) show contour plots of $\theta_+$, $\theta_-$ and $\cos{(\theta_x-\theta_y)}$ corresponding to the denoted states of panel (a). }

214 — 1509.05144

\caption{Schematic visualization of the automaton $\nbw_\text{pump}$ from the proof of Lemma~\ref{thm:buchi_pumpable}. The 3 phases are clearly visible: In the red states {\color{red}$s_{v,x}$} (solid rectangles) the values $(v,x)$ are non-deterministically stored and those states can only be left if there is a change in the value of $r$. The subsequent blue states {\color{blue}$s'_{v,y}$} (dashed rectangles) can only be left in case of a vertex repetition leading to the green state {\color{olive}$s''_z$} (dotted circles) that waits for the next $r'$ change point.}

215 — 1509.05496

\caption{(color online) Standard deviation (a) and skewness (b) of the $Q_k$ distributions versus global packing fraction $\phi$. Each data point represents the average value of the specified moment over several configurations; the associated uncertainty is smaller than the symbol. The main plots emphasize the low $\phi$ behavior, which approaches the same constant value for all preparation protocols shown: infinitely-fast quench ({\color{g} $\times$}), thermal ({\color{b} $\circ$}), randomly placed non-overlapping monodisperse circles ({\color{r} $+$}), and random point patterns with $\phi=0$ (dashed line). Insets show behavior near $\phi_c$. Rescaled distributions of $\phi_c$ are plotted as solid curves, with color indicating preparation protocol from which they were determined. The moments of $Q_k$ have a kink that coincides with the peak in the respective $\phi_c$ distribution. }

216 — 1509.05798

\caption{Step plots of vectors of external $X$ and self citations $Y$ obtained from the IC model, depicted by $\boldsymbol\times$, and Eqs.~\eqref{eq:x_solved} and \eqref{eq:y_solved}, depicted by \textcolor{red}{$\circ$}.}

\caption{Comparison of the $h$-index and its approximations on a Scopus data set. The black continuous line is identity, so ideally all the points should overlie this line. The points depicted with \textcolor{red}{$\circ$} correspond to values given by the $h$-index estimated from the exact solution of the IC simulation with parameters $N_0 = 1, p = 1, q' = 2$, where only the vector of external citations was taken into account, i.e., one that is based on Eq.~\eqref{eq:x_solved}, %+ \eqref{eq:y_solved}, the points marked with \textcolor{blue}{$+$} correspond to the estimate of the $h$-index that is based only on a vector of external citations (Eq.~\eqref{eq:X_approx1}) with parameter $q''=3$, while the points marked with \textcolor{green}{$\bigtriangleup$} correspond to the approximation given by Eq.~\eqref{eq:hirsch_eq} with also $q''=3$. The dotted lines of corresponding color, depicted as \textcolor{red}{$\boldsymbol{---}$}, \textcolor{blue}{$\boldsymbol{-\cdot-\cdot-}$} and \textcolor{green}{$\boldsymbol{\cdot\cdot\cdot\cdot\cdot}$}, are the least squares fit of the $h$-index values and considered approximations, respectively. }

217 — 1509.05906

\caption{(a) Simultaneous measurements of the dynamics of vertical micro-vortex size (\textcolor{blue}{$\bullet$}) and voltage response $\Delta V$ (\textcolor{red}{{\bf--}}), under a constant current density of 10~A/m$^{2}$ across the charge selective membrane. Four characteristic transport regimes are delineated by the dashed lines. \textrm{I}. electric Ohmic conduction without hydrodynamic convection, \textrm{II}. a potential jump with the development of ICP, \textrm{III}. a hydrodynamic convection regime with linearly growing electric resistance and vortex size in time, and IV. a saturated regime with saturated values of both vortex size and electric resistance. (b--d) The corresponding flow field, velocity vectors and vorticity ($\nabla \times \vec{u}$) obtained with a PIV technique at different times; (b) and (c) show the growth and (d) the unsteady nature of the micro-vortices. The vertical arrow in (b--d) indicates the length scale of vortex mixing layer, $L_{mix}$, which initially increases with time. $L_{mix}$ is measured to be 260, 480, and 550~$\mu m$ from (b) to (d), respectively, at different times ($t$) indicated.}

\caption{(a) The growth rate of vortex size (\textcolor{red}{$\bullet$}) and voltage drop $\Delta V$ (\textcolor{black}{$\blacktriangleright$}) during the linear convective Regime \textrm{III} indicated in Fig. 3, under different current densities. (b) The dependence of the average vortex size ($L_{mix}$) (\textcolor{red}{$\blacksquare$}) and speed (\textcolor{black}{$\Diamond$}) on the applied current density across the membrane in the saturated convective regime IV. The error bar shows the standard deviation of the time averaged values, revealing more fluctuations in vortex dynamics at higher currents. (c) The average voltage drops, $\Delta V_{olc}$, over the saturated vortex size in Regime \textrm{IV}, ($\bullet$) obtained from the growth rate data of Regime \textrm{III} (\textcolor{red}{$\blacktriangle$}) and from the data of saturated $\Delta V_{olc}^{IV}$ in Regime \textrm{IV} ($\bullet$). (d) The average conductivity in the mixing layer in the growth regime, $\sigma_{\textrm{III}}$ (\textcolor{red}{$\blacktriangleleft$}), and in the saturated regime $\sigma_{\textrm{IV}}$ ($\bullet$).}

218 — 1509.06239

\caption{\label{fig:fig1} (color online) Dimension of the importance-truncated model space (a) and ground-state energy relative to the core (b) for \isotope[56]{Ni} in the $pf$ valence space as a function of the importance threshold for reference thresholds $C_{\text{min}} = \{1\, $(\symbolcircle[plot1])$,\; 2\, $(\symboldiamondsym[plot2])$,\; 3\, $(\symbolbox[plot3])$\} \times 10^{-4}$ and $T_{\max}=16$ using the \textsc{gxpf1a} interaction. The model space has been constructed for the simultaneous description of the six lowest eigenstates. For the threshold extrapolation we use polynomials of order two and three. The red lines denote the full $m$-scheme dimension and the ground-state energy of the full SM \cite{Horoi2006}. }

\caption{\label{fig:fig2} (color online) Energy-variance extrapolation of the ground-state energy relative to the core for \isotope[56]{Ni} obtained in IT-SM using the \textsc{gxpf1a} interaction. In panel (a) results for different reference thresholds $C_{\text{min}} = \{1\, $(\symbolcircle[plot1])$,\; 2\, $(\symboldiamondsym[plot2])$,\; 3\, $(\symbolbox[plot3])$\} \times 10^{-4}$ for $T_{\max}=16$ are shown. In panel (b) calculations for different truncations $T_{\max} = \{4 $(\symbolcircle[plot1])$,\; 6 $(\symboldiamondsym[plot2])$,\; 8 $(\symbolbox[plot3])$,\; 10 $(\symbolcross[plot4])$\}$ with $C_\text{min} = 1 \times 10^{-4}$ are depicted. }

\caption{\label{fig3}(color online) Energy-variance (a) and threshold (b) extrapolation of the energies of the six lowest natural-parity states of \isotope[56]{Ni} using the \textsc{gxpf1a} interaction with $C_{\text{min}} = \{1\, $(\symbolcircle[plot1])$,\, 2\, $(\symboldiamondsym[plot2])$,\, 3\, $(\symbolbox[plot3])$\} \times 10^{-4}$ and $T_{\max}=16$. For the variance and threshold extrapolations, polynomials of order two and three have been employed, respectively. The red lines show the full SM results extracted from \cite{Horoi2006}. }

\caption{\label{fig:fig5}(color online) Threshold dependence and extrapolation for the quadrupole moment of the $2^+_1$ state (a) and the $B(\text{E2}: 2^+_1 \rightarrow 0^+_1)$ transition strength (b) for \isotope[56]{Ni}. The wave functions have been obtained in an IT-SM calculation using the \textsc{gxpf1a} interaction for $T_{\max}=8$ and the reference thresholds $C_{\text{min}} = \{1\, $(\symbolcircle[plot1])$,\, 2\, $(\symboldiamondsym[plot2])$,\, 3\, $(\symbolbox[plot3])$\} \times 10^{-4}$. The red lines represent the full SM results obtained with the \textsc{Antoine} code \cite{Antoine, Caurier1999, Caurier2005}. }

219 — 1509.06903

\caption{Charts illustrating the layout of each region associated with small-scale extended \gray\emission. From top to bottom: regions around emissivity components E1+E3, E2, and E4. The background is a SHASSA H$\alpha$ smoothed map on a log scale. The green circles are the 1-$\sigma$ extent of the Gaussian emissivity components of the emissivity model, the cyan circles are the 1-$\sigma$ extent of the Gaussian emission components of the analytic model. Overlaid are pulsars (magenta pluses), Wolf-Rayet stars (white diamonds), stars of spectral type B0-3 (blue dots), HMXBs (red dots), and supergiant shells (yellow circles).}

220 — 1509.07341

\caption{\label{fig:1} (color online) a) Time evolution of the average out-of-plane magnetization for different applied current densities %. % and $D = -2\;\mathrm{mJ/m}^2$. %The current varies (variations in steps of $10^{11}\;\mathrm{A/m}^2$). The minimum current to trigger switching, i.e.\the critical current, is highlighted in blue. The green curve indicates the threshold of turbulent behavior (see text). b) Sketch of the magnetization configuration at different stages of the switching process. c)\Mgn\orientation in the center and at the left and right edges. The current induced damping-like torque (represented as effective field\textcolor{mydb}{$\vec H_\text{DL}$}) can only drive the left edge \mgn\into instability, resulting in a nucleation at the left edge. d) Snap-shots of the magnetization configuration showing the reversal from down (black) to up (white) via domain wall nucleation and propagation under an externally applied field of$\mu_0 H=0.1\;$T and a current density of $2.6\;10^{12}\;\mathrm{A/m}^2$. }

221 — 1509.07831

\caption{ \textbf{Learned Deep Point-cloud/Language/Trajectory Embedding Space: } Joint embedding space $h^3$ after the network is fully fine-tuned, visualized in 2d using t-SNE \cite{van2008visualizing} . Inverted triangles represent projected point-cloud/language pairs, circles represent projected trajectories. %with psuedo-classification label by expert which was also used for one of the baselines. The occupancy grid representation (Sec.~\ref{sec:representation}) of object part point-clouds is shown in \textcolor{darkgreen}{green} in \textcolor{blue}{blue} grids. For presentation purpose, `neighbor' cells are not shown. %The occupancy grid representation of object part point-clouds is shown %in \textcolor{darkgreen}{green} in \textcolor{blue}{blue} grids. %Among the two occupancy grids (Sec.~\ref{sec:representation}), %we selected the one that is more visually parsable for each object. The legend at the bottom right shows classifications of object parts by an expert, collected for the purpose of building a baseline. As shown by result of this baseline (object part classifier in Table~\ref{tab:results}), these labels do not necessarily correlate well with the actual manipulation motion. Thus, full separation according to the labels defined in the legend is not optimal and will not occur. % in this figure or Fig.~\ref{fig:task_embed_vis}. These figures are best viewed in color. }

222 — 1509.08071

\caption{\textred{Comparison of RSSI when} \textblue{the back scooter is in the middle lane while the front scooter} \textred{is in different lanes in moving scenarios. The markers represent the mean RSSI while the bounds of the error bars represent 25\% and 75\% percentiles of RSSI.}}

223 — 1509.08679

\caption{ Left: Evolution of the adiabatic energy eigenvalues at the supercell $\Gamma$ point as a function of the position of the projectile along the $\langle 001 \rangle$ channel in Si. The addition of the channeling ion creates a mid-gap ``elevator state'' (red line), the energy of which oscillates as the projectile moves along the channel. The smallest difference between the elevator energy and the valence (conduction) band edge is approximately (less than) $0.2\,$eV. The velocity threshold $E_{\text{th}}'$ calculated using Eq.~(\ref{atompass}) for an energy gap of 0.2 eV is shown in Fig.~\ref{stop_e}. Right: Evolution of the adiabatic energy eigenvalues during a Born-Oppenheimer quantum molecular dynamics simulation in which a Si interstial in a relaxed crystal of 64 initially stationary Si atoms was given an initial velocity of 0.2 $\text{\AA\,fs}^{-1}$ in the $\langle 001 \rangle$ direction. All 65 atoms were allowed to move in response to the interatomic forces generated during the simulation. The initial velocity of 0.2 $\text{\AA\,fs}^{-1}$ is too low for a channeling trajectory to take place, but the elevator-like behavior of the defect states remains apparent. Similar elevator-like behavior was observed when the initial velocity of the projectile was along the $\langle 110 \rangle$ direction. \textcolor{red}{} }

224 — 1509.08701

\caption{Total photoproduction cross section of J/$\psi$ in AuAu collision at $\sqrt{s_{NN}} = 200$ GeV. Comparison with STARLIGHT, Strikman \textit{et al} and Gon\c calves-Machado models, all extracted from \cite{PHENIX2009} as well as the experimental data. The lines (\textendash \textperiodcentered) delimit the statistical uncertainties in the calculations.}

225 — 1510.00477

\caption{Ranking of photos according to our model's visual realism prediction. The color of image border encodes the human annotation: {\color{green}green}: realistic composites; {\color{red}red}: unrealistic composites; {\color{blue}blue}: natural photos. The different rows contain composites corresponding to different rank percentiles of scores predicted with \textit{RealismCNN + SVM}.}

226 — 1510.00494

\caption{Phase diagrams with the evolution of contours from $\protect\mu % =10^{3}$ GeV (\textbf{Background-Left}) up to $\protect\mu =10^{19}$ GeV (\textbf{Background-Right}) in the $\Delta S_{1}^{2}$ versus $\protect\lambda _{1}\left( m_{Z}\right)$ plane. Here $0\leq\protect\lambda _{2}(m_{Z})\leq0.25$ and $-0.25\leq\protect% \lambda _{3,4}(m_{Z})\leq0.25$, starting with $\protect\lambda _{3,4}(m_{Z})=% \protect\lambda _{5}(m_{Z})/2$ and $\protect\lambda _{34}(m_{Z})=\vert \protect\lambda_{2}(m_{Z})\vert$. Red lines are the remaining contours between $% \protect\mu =10^{3}$ and $10^{19}$ GeV. Dashed line indicates the experimental value for the ratio in $\protect\lambda _{1}(m_{Z})$ for a Higgs with a mass near to $125$ GeV \cite{Aad}. Gray curve encloses region compatible with the strongest unitarity bound given by the eigenvalue $\boldsymbol{\Lambda}_{00}^{even+}$.} \label{fig:L4L1} \end{figure} \begin{figure}[tph] \centering\includegraphics[scale=0.18]{L1L5Z3.pdf} % \includegraphics[scale=0.18]{L1L5Z19.pdf} \vspace{0.6cm} \caption{Phase diagrams with the evolution of contours from $\protect\mu % =10^{3}$ GeV (\textbf{Background-Left}) up to $\protect\mu =10^{19}$ GeV (\textbf{Background-Right}) in the $\Delta S_{0}^{2}$ versus $\protect% \lambda _{1}\left(m_{Z}\right)$ plane. Here $0\leq\protect\lambda _{2}(m_{Z})\leq0.25$ and $-0.25\leq\protect% \lambda _{3,4}(m_{Z})\leq0.25$, starting with $\protect\lambda _{3,4}(m_{Z})=% \protect\lambda _{5}(m_{Z})/2$ and $\protect\lambda _{34}(m_{Z})=\vert \protect\lambda_{2}(m_{Z})\vert$. Red lines are the remaining contours between $% \protect\mu =10^{3}$ and $10^{19}$ GeV. Dashed line indicates the experimental value for the ratio in $\protect\lambda _{1}(m_{Z})$ for a Higgs with a mass near to $125$ GeV \cite{Aad}. Gray curve encloses the region compatible with the strongest unitarity bound given by the eigenvalue $\boldsymbol{\Lambda}_{00}^{even+}$.}

227 — 1510.00635

\caption{\label{fig:HPGS01}\HESS\Galactic Plane Survey region, flux map and exposure map (from top to bottom). The all-sky image on the top panel shows a Planck CO Map with\LAT\identified Galactic 1FHL sources (triangles) and the 15 known Galactic TeV sources (white stars) outside the HGPS region. The HEGRA Galactic Plane Survey~\cite{hegra-survey} and the VERITAS Cygnus survey~\cite{2009arXiv0912.4492W,veritas-cygnus-icrc2015} regions are illustrated in blue and green, respectively. From~\cite{hess-hgps-icrc2015}.}

\caption{\label{fig:RXJ1713}{\bf Left}: \HESS VHE \gray\image of RX~J1713.7-3946.{\bf Right}: Map of magnetic field under a leptonic scenario~\cite{hess-rxj1713-icrc2015}.}

\caption{\label{fig:RXJ1713_Profiles}{\bf Left}: Quadrants used in the \HESS VHE \gray\image of RX~J1713.7-3946 to investigate possible particle escape.{\bf Right}: Radial profile of \gray\(black) and X-ray (red) emission in quadrant number 3~\cite{hess-rxj1713-icrc2015}.}

\caption{\label{tab:Binaries}Orbital period and companion star mass and temperature of the five \gray binary systems known so far. From ~\cite{2013A&ARv..21...64D} and references therein.}

\caption{\label{fig:BinariesGeometry}The two main scenarios for \gray\emission from binaries.{\bf Left}: in the pulsar wind scenario, the variable emission arises from the interaction of the pulsar wind with the strong stellar wind of the companion star. {\bf Right}: in the microquasar scenario, the emission is powered by the accretion of the massive star onto the compact object (black hole or neutron star) giving rise to relativistic jets. Reproduced from~\cite{2013A&ARv..21...64D}.}

228 — 1510.01148

\caption{\footnotesize Performance in terms of "center location error" (CLE) in pixels. \textcolor[rgb]{1,0,0}{Red} and \textcolor[rgb]{0,0,1}{blue} colors indicate the best and 2nd best performance, respectively.}

\caption{\footnotesize Performance in terms of "overlap rate" $e$ (in pixels). \textcolor[rgb]{1,0,0}{Red} and \textcolor[rgb]{0,0,1}{blue} colors indicate the best and 2nd best performance, respectively.}

229 — 1510.01223

\caption{ Characterization of the trajectories as a function of the orbital angular momentum $l$ and the spin $s$, for $e =1=M$. The trajectories are run inward from large $r$ to small $r$ and it is tested for each step whether, a radial turning point, a horizon crossing, or a divergent center of mass energy (divergence radius) occurs. The color coding is:\\ \textcolor{purple}{purple} $=$ turning point first, {\bf{black}} $=$ horizon first, \textcolor{yellow}{yellow} $=$ divergence radius first. White points correspond to trajectories which have the radial turning point and the divergence radius too close together to be distinguished by the chosen numerical step size ($\delta r/r=32.000$).}

230 — 1510.01431

\caption{Example images with neural, positive (\textcolor{green}{green}) and negative (\textcolor{red}{red}) captions, by crowd workers in MSCOCO dataset~\protect\cite{chen2015microsoft} and this work (Section~\protect\ref{sec:mturk}). }

\caption{(a) One example image with both \colorbox{green!20}{positive} and \colorbox{red!20}{negative} captions written by AMT workers. (b) Summary of quality validation for sentiment captions. The rows are MSCOCO~\protect\shortcite{chen2015microsoft}, and captions with {\sc Pos}itive and {\sc Neg}ative sentiments, respectively. {\em Descriptiveness $\pm$ standard deviation} is rated as 1--4 and averaged across different AMT workers, higher is better. The {\em Correct sentiment} column records the number of captions receiving 3, 2, 1, 0 votes for having a sentiment that matches the image, from three different AMT workers. }

\caption{Example results from sentiment caption generation. Columns a+b: \textcolor{green!80}{positive} captions; columns c+d: \textcolor{red!50}{negative} captions. Background color indicate the probability of the switching variable $\gamma^1_t=p(s_t|\cdot)$: \colorbox{green}{da}\colorbox{red}{rk} if $\gamma^1_t \ge 0.75$; \colorbox{green!60}{med}\colorbox{red!60}{ium} if $\gamma^1_t \ge 0.5$; \colorbox{green!30}{lig}\colorbox{red!30}{ht} if $\gamma^1_t \ge 0.25$. Row 1 and 2 contain generally successful examples. Row 3 contains examples with various amounts of error in either semantics or sentiment, at times with amusing effects. See Section~\ref{sec:exp} for discussions. }

231 — 1510.01787

\caption{ The strange form factors $G_E^{(s)}$ and $G_M^{(s)}$~\protect\cite{Shanahan:2014tja}. The \textcolor{red}{red} and \textcolor{orange(colorwheel)}{orange} ellipses are the results of LQCD calculations at $Q^2=0.26~{\rm GeV}^2$ and $Q^2=0.17~{\rm GeV}^2$, respectively, while the \textcolor{darkpowderblue}{blue} and \textcolor{forestgreen(web)}{green} ellipses show the experimental constraints at $Q^2=0.23~{\rm GeV}^2$. [Figure reproduced with permission from Phiala Shanahan.] }

232 — 1510.01893

\caption{ (a) TeV \gray\radial surface brightness profile of sector ``region 3." The\hess\flux data and the{\sl XMM-Newton} X-ray brightness profile convolved with the \hess\PSF of$3'$ (gray line) are adopted from \citet{deNaurois2015}. The radial profile of ``region 3" is fitted with a two-zone emission (in black), which is the sum of the inner (in green) and outer (in blue) components. The solid and dashed lines represent the profiles in \caseA\and\caseB, respectively. All of the fitting profile lines have been convolved with the \hess\PSF of$3'$. (b) TeV \gray\radial surface brightness profile for the entire remnant. The\hess\flux data are adopted from\citet{J1713.HESS.2006}. The meanings of all lines convolved with the \hess\PSF of$5'$ are the same as those in (a). }

233 — 1510.02251

\caption{N$_2$D$^+$(2--1) integrated intensity map overlaid on the N$_2$H$^+$(1--0) and NH$_3$(1, 1) integrated intensity maps. The contour level parameters of N$_2$D$^+$(2--1) integrated intensity map are the same as in Fig.~\ref{fig_son2dp}. The thin black contour in each panel indicates the 3$\sigma$ level of the N$_2$H$^+$(1--0) and NH$_3$(1, 1) integrated intensity emission. The $\sigma$ values and integrated velocity range are the same as in Figs.~\ref{fig_son2dp} and \ref{fig_lines}, respectively. The star marks the UCH{\sc ii} region. The dotted contours are the negative features due to the missing flux with the same contour levels as the positive ones in each panel. The synthesized beam of N$_2$D$^+$(2--1) is shown in the bottom left corner of each panel, and the one of N$_2$H$^+$(1--0) and NH$_3$(1, 1) is shown in the bottom right corner of \textcolor{red}{respective} panel.}

\caption{Molecular line velocity maps overlaid with 2~mm, 3~mm continuum map and the NH$_3$(1, 1) integrated intensity map. The contour level parameters are the same as the ones in Figs.~\ref{fig_mmcont} and \ref{fig_lines}. The dashed arrows in the {\it top} panels mark the PV-cuts used for PV diagrams in Fig.~\ref{fig_pv}. The star marks the UCH{\sc ii} region. The cross in the {top right} panel indicateks the reference position in CH$_3$OH PV diagram in Fig.~\ref{fig_pv}. The circles indicate the area where the average spectra are extracted and shown in Fig.~\ref{fig_chspec}. The synthesized beam of the velocity maps is shown in the bottom left corner of each panel, and the one of the contours is shown in the bottom right corner of \textcolor{red}{respect} panel.}

\caption{Molecular line-width maps overlaid with the 2~mm, 3~mm continuum map and NH$_3$(1, 1) integrated intensity map. The contour level parameters are the same as the ones in Figs.~\ref{fig_mmcont} and \ref{fig_lines}. The dashed arrows in the {\it top} panels point to the PV-cuts used for PV diagrams in Fig.~\ref{fig_pv}. The star indicates the UCH{\sc ii} region. The cross in the {top right} panel marks the reference position in CH$_3$OH PV diagram in Fig.~\ref{fig_pv}. The circles show where the average spectra are extracted and shown in Fig.~\ref{fig_chspec}. The synthesized beam of the line-width maps is shown in the bottom left corner of each panel, and the one of the contours is shown in the bottom right corner of \textcolor{red}{respect }panel.}

\caption{Position-velocity (PV) diagrams of C$_2$H and CH$_3$OH with a velocity resolution of 0.42~km~s$^{-1}$. The PV cuts are shown in the \textcolor{red}{respect} velocity map (Fig.~\ref{fig_vel}). The contour levels are from 10\% to 90\% from the peak emission (from left to right, 0.74, 0.69, and 0.88~Jy~beam$^{-1}$, respectively) in steps of 10\%. The offsets refer to the distance along the cuts from the reference position, which is, from left to right, the dust continuum peak of MM1, the cross position of pv-1 and pv-2 in Fig.~\ref{fig_vel}, and the cross in the CH$_3$OH panel in Fig.~\ref{fig_vel}, respectively. The $v_{\rm LSR}$ at --18.3~km~s$^{-1}$ and the reference position are indicated by vertical and horizontal dashed lines. }

\caption{Convolved mm continuum maps. {\it Left:} The 3~mm continuum map (red) overlaid on the 2~mm continuum map (gray scale and black contours), where both maps were produced with the same {\it uv-}range and same beam size. {\it Right:} The 1.3~mm continuum map (red) overlaid on the 2~mm continuum map (gray scale and black contours), where both maps were produced with the same {\it uv-}range and same beam size. The triangles mark the 2~mm cores, and the squares the 1.3~mm cores detected by \citet{palau2013}. The star indicates the UCH{\sc ii} region VLA1. All the contours in both panels start at 4$\sigma$ and increase in steps of 2$\sigma$. The $\sigma$ value in mJy beam$^{-1}$ and the {\it uv-}range are shown in \textcolor{red}{respect} panel. The dotted contours are the negative features due to the missing flux with the same contour levels as the positive ones in each panel. The synthesized beam is shown in the bottom right corner of each panel.}

234 — 1510.02521

\caption{{The cells $X \cup X'$ (\textcolor{red}{bold face}) and $Y \cup Y'$ (\textcolor{blue}{underlined bold face}) are shown for $L_{\Bbb{Z}_{12}}$ and for $L_{\Bbb{Z}_{14}}$.}}

235 — 1510.02602

\caption{\label{li6bi_sigedep}(Color online) \textcolor{red}{Integrated cross sections for $^{6}$Li on $^{209}$Bi as a function of the incident laboratory energy. The crosses are the EBU results, the stars the complete fusion (CF) experimental data, taken from \cite{Das04}. The open squares are the NEB calculations treating $\alpha$ as spectator, deuteron absorption, obtained with the FR-DWBA closed-form formula, whereas the diamonds are the NEB calculations treating deuteron as spectator, $\alpha$ absorption. The solid line is the sum of CF, NEB-$\alpha$, NEB-deuteron, and EBU. The solid circles are the reaction cross sections, obtained with the CDCC calculations. The arrow indicates the nominal position of the Coulomb barrier.}}

236 — 1510.02668

\caption{For each configuration in the monogaussian case, one realization of $F$ and the associated $y$ (\hdashrule[0.5ex]{1cm}{0.5pt}{}), $s_1(x)$ and $s_2(x)$ ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}). The coding function is constant for (a) and (c), and variable for (b) and (d). The covariance model is $C_1$ for (a) and (b), and $C_2$ for (c) and (d).}

\caption{Simulation results in the monogaussian cases with covariance $C_1$ (left) and $C_2$ (right), in the constant coding function case (up) and in the varying case (middle). The down figures stand for the Gaussian case. Actual model ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), empirical variogram of ten arbitrary simulations ({\color{gray}\hdashrule[0.5ex]{1cm}{0.1pt}{}}), average of the empirical variograms over all the simulations (\hdashrule[0.5ex][x]{1cm}{1pt}{2.5mm 1mm 1pt 1mm}), 25th and 75th percentiles ( {\color{blue}\hdashrule[0.5ex]{1.2cm}{1pt}{1pt 1mm}}), 5th and 95th percentiles ( {\color{blue}\hdashrule[0.5ex][x]{1.2cm}{1pt}{1.5mm}}).}

\caption{ Actual model ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), empirical ominidirectional variogram of ten arbitrary simulations ({\color{gray}\hdashrule[0.5ex]{1cm}{0.1pt}{}}), average of the empirical variograms over all the simulations (\hdashrule[0.5ex][x]{1cm}{1pt}{2.5mm 1mm 1pt 1mm}), 25th and 75th percentiles ( {\color{blue}\hdashrule[0.5ex]{1.2cm}{1pt}{1pt 1mm}}), 5th and 95th percentiles ( {\color{blue}\hdashrule[0.5ex][x]{1.2cm}{1pt}{1.5mm}}).}

\caption{ Experimental variograms of the underlying GRFs obtained by PL and fitted model ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}) . Up : vertical; down: horizontal (omnidirectional in the horizontal plane). From left to right: $q=1$, $q=2$ and first GRF, $q=2$ and second GRF. }

\caption{Vertical indicators simple and cross-variograms (from up to bottom and left to right: orange, black and green): computed by PL from the categorical data ({\color{black}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), computed by equation (\ref{varioexpindic}) from the monogaussian model ({\color{green}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), from the plurigaussian model ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}).}

\caption{Horizontal indicators simple and cross-variograms (from up to bottom and left to right: orange, black and green) : computed by PL from the categorical data ({\color{black}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), computed by equation (\ref{varioexpindic}) from the monogaussian model ({\color{green}\hdashrule[0.5ex]{1cm}{0.5pt}{}}), from the plurigaussian model ({\color{red}\hdashrule[0.5ex]{1cm}{0.5pt}{}}).}

237 — 1510.02966

\caption{\textbf{Demonstration of the superposition rule arising from the flow model.} Distribution of force on the perimeter of two long moving objects as calculated directly from continuum simulations of our theory (\color{blue}$\to$\color{black}), compared to predictions from RFT using theoretical RFP's from figure (2) (\color{red}$\to$\color{black}). Net resistive force shown at the center of the object.}

\caption{\textbf{Validity check for superposition in 3D.} The drag force, $F_x$, acting on a rightward moving submerged `V' intruder ($\times$), and the components of lift, $F_z$ (\color{red} $\bigtriangledown$\color{black}), and lateral force, $F_y$ (\color{blue} $\triangle$\color{black}), compared with superposition values (dashed) at various orientation angles, $\varphi$, for vertical (top row) and horizontal (bottom row) intruder alignments in granular flow (left, $\rho=4 $g/cm$^3$, $\mu=$0.4, depth=0.6m) and viscous fluid (right, $\eta v$=15N/m). Intruder consists of two square plates of side length 20cm. Grey regions indicate orientations where the plates are not both on the leading surface of the V.}

\caption{\textbf{Evaluation of superposition for a flat intruder.} The drag force, $F_x$, acting on a rightward moving submerged flat intruder ($\times$), and the components of lift, $F_z$ (\color{red} $\bigtriangledown$\color{black}), and lateral force, $F_y$ (\color{blue} $\triangle$\color{black}), compared with superposition values (dashed) at various orientation angles, $\varphi$, for horizontal intruder alignments in granular flow (left, $\rho_c=4 $g/cm$^3$, $\mu=$0.4, depth=0.6m) and viscous fluid (right, $\eta v$=15N/m). Intruder consists of two square plates of side length 20cm.}

238 — 1510.03981

\caption{ \textbf{ Differential temperature response and the quantum of thermal conductance. } \textbf{a}, Measured (dots) and simulated (dashed lines) temperatures of Island~A as functions of the temperature of Island~B for the indicated phonon bath temperatures in Sample A1. The results for the control sample at 100 mK bath temperature are shown for comparison. \textbf{b}, Differential temperature response (\mbox{\boldmath${\pmb\circ}$}) from (\textbf{a}) at the lowest $T_\textrm{B}$ for each bath temperature. The experimental uncertainty is of the order of the marker size. For comparison, we also show the corresponding experimental data for the control sample (\textcolor{magenta_1}{\mbox{\boldmath${\pmb \lozenge}$}}), for Sample A2 (\textcolor{blue}{\boldmath${\pmb \triangleright}$}), and for A3 (\textcolor{green}{\boldmath${\pmb \medsquare}$}). The solid black line shows the prediction of the full thermal model of Fig.~\ref{fig:thermal_model}a. The dashed lines are calculated with 80~\% (bottom) and 115~\% (top) of the quantum of thermal conductance indicating the sensitivity of the results to the photonic heat conduction. The solid red lines are calculated with the simplified thermal model [Eq.~\eqref{eq:simplified_theory}] for Sample A1 without any free parameters: we take into account only quantum-limited photonic heat conduction and electron--phonon coupling with literature values \cite{Giazotto} for the constant of Cu, $\Sigma_\textrm{N} = 2\times10^9$~WK$^{-5}$m$^{-3}$ (right), and $\Sigma_\textrm{N} = 4\times10^9$~WK$^{-5}$m$^{-3}$ (left). }

239 — 1510.04198

\caption{Optimal beam waist radius $\sigma$ [left] and the intensity coupled into the fiber relative to the incident [\textcolor{blue}{${P_{\text{ein}}}/{P_{\text{in}}}$}] and the mode purity of $LP_{01}$ [\textcolor{cyan!80!white}{$P_{11}/P_{\text{ein}}$}] over the free space propagation distance $\zeta$ using phase plate \protect \phasePl{1}.}

\caption{Optimal beam waist radius $\sigma$ [left] and the into the fiber coupled intensity relative to the incident [\textcolor{blue}{${P_{\text{ein}}}/{P_{\text{in}}}$}] and the mode purity of $LP_{02}$ [\textcolor{cyan!80!white}{$P_{21}/P_{\text{ein}}$}] over the free space propagation distance $\zeta$ using phase plate \protect \phasePl{2}.}

240 — 1510.04862

\caption{\red{For a discovered TRO (tap): multiple users enable modelling the object's position $\Phi_k$ (top-left) learning varying views in the appearance model $A_k$ (top-right) %; building an approximate 3D model $M_k$ (bottom-left) and gathering different usage snippets $U_k$ that show interactions with the same object (bottom).}}

\caption{\red{Recall, precision and F1-score} results for discovering TROs using different features, clustering methods, with/without attention and sliding window for the proposed offline and online TRO discovery methods.}

241 — 1510.04908

\caption{\includegraphics[height=0.0125\textheight]{images/icons/bottle.png} 4097 $\rightarrow$ 340}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/sofa.png} 4446 $\rightarrow$ 2081}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/bus.png} 262 $\rightarrow$ 130}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/aeroplane.png} 456 $\rightarrow$ 539}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/sofa.png} 372 $\rightarrow$ 128}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/table.png} 1443 $\rightarrow$ 404}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/horse.png} 120 $\rightarrow$ 80}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/tv.png} 3748 $\rightarrow$ 4527}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/boat.png} 30 $\rightarrow$ 12}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/cat.png} 436 $\rightarrow$ 309}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/sheep.png} 796 $\rightarrow$ 242}

\caption{\includegraphics[height=0.0125\textheight]{images/icons/sheep.png} 479 $\rightarrow$ 560}

242 — 1510.04913

\caption{The left and center panels are the ground-state phase diagrams for the zero magnetic field case with $\tilde \Jm = 0.0$ and $\tilde \Jm = 0.3$, respectively. The TLL-dimer boundary (\opencircle) is of the BKT type, while other boundaries (\opensquare \and\fullcircle) are of the first order. The right panel shows the spin configurations for (a) the ferrimagnetic state and (b) the dimerized state, where ellipses denote the singlet pairs. The dimerized state is doubly degenerate. }

\caption{The left and center panels are the ground-state phase diagrams for the $M=\Ms/3$ case with $\tilde \Jm = 0.0$ and $\tilde \Jm = 0.3$, respectively. The boundary (\opensquare) between the ferromagnetic phase and the no $\Ms/3$ plateau phase is of the first order, and that (\fullsquare) between the $\Ms/3$ plateau phase to the no $\Ms/3$ plateau phase is of the BKT type. The plateau mechanism changing line ($\times$) is of the Gaussian type. The right panel shows the spin configurations for (a) the type A plateau and (b) the type B plateau, where ellipses denote the singlet pairs. The spin configuration of the type A plateau is the same as that of the ferrimagnetic state under zero magnetic field in the right panel of Fig.2. }

\caption{Ground-state phase diagrams for the $M=(2/3)\Ms$ case with $\tilde \Jm = 0.0$ (left), $\tilde \Jm = 0.1$ (center) and $\tilde \Jm = 0.3$ (right). The boundaries (\opendiamond) between the plateau phase and the no plateau phase are of the BKT type. }

243 — 1510.05186

\caption{ Left: % The contour where the first contribution(Eq.\eqref{Boltzmann1}) can explain the observed baryon asymmetry. % Here we take \blue{$\Lambda_\text{inf}=10^{15}$GeV,} $\sum_j \lambda_{1,jj}\mathrm{Im}(\lambda_{2,1j})=2$ and $m_\nu=0.1\eV$. % %\blue{ %\magenta{(} %Notice that the initial value of $a$, $a(t=t_\text{end})$, is not important because we can always choose the normalization of $a$ freely. %\magenta{)} %} \magenta{ Note that %the initial value of $a$, $a(t=t_\text{end})$, is not important because the normalization of $a$ can be chosen freely, and we take $a(t=t_\text{end})=1$. We numerically confirmed that the asymmetry is mainly produced at the time where the reheating is just completed, $t\sim\Gamma_\text{inf}^{-1}$, and therefore the result is insensitive to the value of $\Lambda_\text{inf}$ as long as $T_R\lesssim\Lambda_\text{inf}$. } % The solid and dashed lines correspond to $\Lambda_2=10^{14}\GeV$ and $\Lambda_2=10^{15}\GeV$, respectively. % The red and blue lines represent $\mathrm{Br}=1$ and $0.1$. % In the blue region, because $T_R>m_\infla$, we need more careful study of the reheating, which is beyond scope of this paper. % Right: % Same as left figure, but considering the second contribution(Eq.\eqref{Boltzmann2}), taking $\sum \mathrm{Im}(y_{1i}^* y_{kl} \lambda_{2,1ikl})/|y_{1i}|^2=1$ and assuming $\lambda_{1,11}^2\gg\lambda_{1,22}^2, \lambda_{1,33}^2$. }

\caption{The diagram contributing to the lepton asymmetry. \red{The solid and dotted lines represent the lepton and Higgs doublet, respectively.} }

\caption{The decay of an inflaton. % The interference between the tree and the one-loop diagram induces the CP asymmetry. \red{ Here the dotted line corresponds to the inflaton. $i, j$ are the flavor indexes. % } }

\caption{ \red{ The washout process of lepton flavor asymmetry. % $E$ is the right handed charged lepton field. } }

244 — 1510.05625

\caption{{\bf The critical fraction of frozen chains $c^\dag$ at which all the non-explicitly frozen chains are usually caged by threadings decreases with ring length.} The mean squared displacement of the centre of mass ($g_3(t)$) averaged over only the non-frozen fraction ($1-c$) of rings plotted against time for different values of the frozen fraction $c$ ranging from $c=0$ (free solution) to $c \geq c^\dag$ at which all rings in the un-frozen fraction have become permanently caged. \textbf{(a)} $M=256$, \textbf{(b)} $M=512$ \textbf{(c)}$M=1024$, \textbf{(d)} $M=1512$ and \textbf{(e)} $M=2048$. The thick horizontal dashed lines indicate the diameter of the chains. \textbf{(f)} The value of $D_{\rm eff}$ (eq.~\eqref{eq:Deff}) normalized by $D_0\equiv D_{\rm eff}(c=0)$ is shown \red{in (main) linear-linear and (inset) linear-log scale to appreciate the nearly exponential dependence}. The critical value $c^\dagger$ is defined as the one for which $D_{\rm eff}(c^\dagger)/D_0$ is below a critical threshold that we here set to $0.02$ (dashed line). }

\caption{(a) \red{Ensemble average of the radius of gyration $\langle R_g \rangle \equiv \langle R^2_g \rangle^{1/2}$ and plotted against rings' contour length $s$ (main) and total length $M$ (inset) both in units of the bead size. As reported in the literature, the rings show a collapsed behaviour for large $M$}. (b) Static scattering function $S_1(q)$ plotted against $qR_g$ and normalised by the length of the rings $M$. This quantity indicates a complex arrangement of the rings internal structure, which does not seem to follow a unique fractal dimension at all lengths. }

\caption{(a) The number of surface monomers $m_s$ shows a scaling $m_s \sim M^{\beta_c}$ with $\beta_c \simeq 0.99 $ for the longest chains. (b) The number of contiguous chains $n_c$ and number of neighbouring chains $n_n$ is shown. While their scaling behaviour is similar, the average number of chains in contact at any time is systematically larger than the number of chains whose centre of mass is closer than $2R_g$. This may imply large fluctuations in the rings conformations that bring distant chains in contact with one another. \red{In (b) we show two curves with exponents $0.33$ and $0.27$ as a guide for the eye but refer to the literature and further studies for more precise estimates.}}

\caption{(a) The contact probability function $P_c(m)$ (defined in eq.~\eqref{eq:Pc}) shows a scaling behaviour $m^{-\gamma}$ with $\gamma$ slightly dependent on the choice of the cut-off $a$ but compatible with previous findings~\cite{Halverson2011}. (b) Pair correlation function $g(r)$ for the rings centre of mass as defined in eq.~\eqref{eq:si_gr}. The peak position imply that the coils are largely overlapping. \red{In the inset we report $g(r)$ normalised by the ideal pair distribution function $g_I(r)=4\pi\rho_c r^2$, where $\rho_c \equiv N/L^3$ is the coils' density. The coils behave as ultra-soft sphere with large inter-penetrations.} }

\caption{\textbf{(a)} $g_3(t)$: Mean square displacement (MSD) of the rings centre of mass. The faint horizontal lines represent the square radius of gyration $\langle R^2_g (t)\rangle$. \textbf{(b)} $g_3(t)/R_g^2$: MSD of the rings centre of mass divided by their squared gyration radius. The solid horizontal line in \textbf{(b)} marks the value $g_3(t)/R_g^2=1$ at which the rings have, on average, travelled once their own size. \textbf{(c)} and \textbf{(d)} Report the behaviour of the diffusion coefficient of the centre of mass of the rings and the relaxation time $\tau_R$ (see eq.~\eqref{eq:tau_relax}), respectively. \red{By fitting the last three data-points in order to obtain the asymptotic values of $D_{CM}$ for large $M$ we obtain $-2 \pm 0.1$. By considering all data point we obtain a value for the exponent of $D_{CM}$ of around $1.8 \pm 0.1$.} }

\caption{%\textbf{(a)} Persistent contiguous chains $\varphi_{nc}$ as defined in eq.~\eqref{eq:PersNeigh}; solid lines represent stretched exponentials $\varphi \sim \exp{-\left(t/\tau_{nc}\right)^\beta_{nc}}$, while the dashed line on the right is guide for the eye and indicates a power law of $t^{-0.75}$. $\beta\simeq 0.5$ and the dashed line a power law $t^{-3/2}$ \textbf{(a)} Values of the relaxation times $\tau_{nc}$ \red{and $T_{\rm nc}$} for the contiguous correlation function $\varphi_{nc}(t)$ reported in Fig.~2 of the main text.%\ref{fig:Main_PersNeigh}. We observe an exponential increase of the typical exchange time for large $M$. \textbf{(b)} Value of the stretching exponent $\beta_{nc}$, ranging from near unity for $M=256$ to around $1/2$ for $M=2048$.}

245 — 1510.06786

\caption{The observed scores computed by \dist\(in \protect \hwplotgreen) for \texttt{buffalo} and \texttt{hand} when analyzing regional differences between New York and USA overall. The histogram shows the distribution of scores under the null model. The $98\%$ confidence intervals of the score under null model are shown in \protect \hwplotred. The observed score for \texttt{hand} lies well within the confidence interval and hence is not a statistically significant change. In contrast, the score for \texttt{buffalo} is far outside the confidence interval for the null distribution indicating a statistically significant change.}

246 — 1510.07387

\caption{The peak antenna temperature variations of the SiO $v$ = 1 and $v$ = 2, $J$ = 1--0 masers (black and open circles) are plotted on the optical light curve of V407 Cyg from AAVSO ({\color{red} red circles}) during our monitoring observations from March 2, 2010 to June 5, 2014. Optical maxima are indicated by the dashed vertical lines shown in black. The arrows indicate the upper limits of the SiO masers. The dashed vertical line shown in {\color{red} red} indicates the nova outburst epoch of March 10, 2010. The X-axis denotes Julian dates. \label{fig:jkasfig2}}

247 — 1510.07486

\caption{M\"ossbauer spectra for Fe$_{1+y}$Te and fits (black curves) for $y=$~0.06 (a), 0.13 (b), and 0.15 (c). Corresponding $T$-dependence of the magnetic hyperfine field $B_{\rm hf}$ for $y=$~0.06 (d), 0.13 (e), and 0.15 (f). Black circles and dashed lines correspond to the values obtained for pattern 2 ($cf.$ text) while red lines and squares correspond to the average value of the magnetic hyperfine field distribution, which were obtained by \red{the} maximum entropy method (the error bars show standard \red{deviations}). Lines are \red{guides to the eye}.}

\caption{(a)--(c) Magnetic states described by the phenomenological Dzyaloshinskii model, Eq.~(\ref{f}). (The staggered atomic spin-structure in $c$-direction is represented by one sublattice only for clarity.) (a) A one-dimensional cycloidal spiral transverse (shown arbitrarily in $c$ direction) to the propagating incommensurable spin-density wave along $b$ (upper half) represents the LRO state. Lower half shows the commensurate collinear SDW-state. (b) Two-dimensional skyrmionic texture with a double-twisted core in the $ac$-plane. (c) Three-dimensional solitonic state. (d) Profiles of one-dimensional states of modulated, $s_b \propto |S|\,\exp(i\phi)$, or twisted form, $(s_b,s_c) \propto |S|\,(\cos(\theta),\sin(\theta))$: (i) Commensurate states with $|S|$ solitons below the lock-in transition. (ii) Incommensurate LRO. (iii) Isolated amplitude-modulated solitons that arise as fluctuations in the precursor state. (e) Schematic phase diagram for Fe$_{1+y}$Te. A spin-liquid state of solitons like those shown in (b) and (c) with amplitude-modulation (d(iii)) precede the incommensurate SDW-like state. Triangles mark experimental transition temperatures for the onset of the precursor and of $T_N$. At $T_l$ a lock-in \red{into} a commensurate state occurs. Point 'L' is a \red{tricritical} Lifshitz point of the spin system associated with structural phase transitions. }

248 — 1510.07595

\caption{A passive tensegrity structure modeled according to the elbow by Dr. Graham Scarr. \textcolor{red}{\emph{Is this figure from another paper? If so, it is generally poor form to copy and past figures from other papers (specially without citing the figure directly). Usually one just references the paper itself. When you really need to show another author's figure, recreate it.}}}

249 — 1510.07867

\caption{Preferences between clusters of users. The color of the arrow indicates how much the men's cluster likes (\textcolor{ForestGreen}{green}) or dislikes (\textcolor{red}{red}) the women's cluster on average. }

250 — 1510.08238

\caption{Pictures of the two boards composing the TRB mounted on the NAT\textregistered \MCH. The main TRB board is the top board in the left picture. The extension PCB providing the SMA connectivity with the front panel is not mounted.}

251 — 1510.09221

\caption{Simulation results of various initial conditions which give rise to nonexplosive solutions (indicated by {\color{blue} $\ast$}) and explosive solutions (indicated by {\color{red} $\diamondsuit$}) of our $\mathbb{C}^2$-valued coupled system (\ref{eq:1}), with $\beta=\nu=1$. The results on the left panel are for the system without added Brownian noise, while the results on the right panel are with added Brownian noise of the form (\ref{eq:14}).}

252 — 1511.00006

\caption{ Dust maps of M31. \emph{Top} - the mean dust extinction map from \dalcanton{} which we use throughout this paper. The map shows that the dust extinction is highest in the 10~kpc ring region, with lower extinction values in other star-forming regions. \emph{Bottom} - map of the fraction of RGB stars that are reddened, \fred{} from \dalcanton{}. The \fred{} parameter is important in understanding structure caused by dust; e.g., at a fixed $A_V$ increasing \fred{} means more stars will be extincted. The final \dalcanton{} dust parameter, \avsig{}, is roughly 0.3 everywhere and is not shown here because its effect on our results is negligible. }

\caption{ The metallicity distribution function (MDF) of our RGB star sample using a flat fiducial age of 4~Gyr. The cyan MDF uses this paper's RGB selection. In purple, we show only RGB stars from low extinction regions within the PHAT survey (\avfred{} $< 0.25$ mag; see Section~\ref{sec:dust-sources}). The median metallicity for both histograms is $\sim$-0.11~dex (shown as a vertical line). We note this MDF is not corrected for the systematic biases in the metallicity discussed in Section \ref{sec:artificial-star-tests} and \ref{sec:dust-impact}, however, after correction, the MDFs look qualitatively very similar. }

\caption{ Simulated change in median metallicity ($\Delta$[M/H]) for each spatial bin in the data. The simulation uses a fiducial CMD taken from a low-dust region and applies reddening to individual stars using the dust parameters (\fred{},\avdal{}, and \avsig{}) from the \dalcanton{} dust maps for every spatial bin. The solid lines show changes in median metallicity at constant \fred{} values of 0.1, 0.2, 0.3, and 0.4. Along one of these solid lines (and assuming \avsig{}=0.3). The dust increases the median metallicity until \avdal{}$\approx$0.5 because the reddened stars receive an overestimated metallicity. The median metallicity decreases for high extinction when the reddened stars are moved beyond the edge of our RGB box and no longer affect the measurement. The points below the dashed line are low extinction regions with \avfred{} $< 0.25$. These are the regions used in our metallicity gradient measurement. Their $\Delta$[M/H] is $<0.12$~dex, with typical values of 0.03~dex. Details on the dust simulations are given in Section~\ref{sec:dust-impact}; a map of $\Delta$[M/H] is shown in the bottom panel of Figure~\ref{fig:uncorrected-med-metal}. }

\caption[width=0.9\paperwidth]{ {\em Top Panel}~-- Map of the RGB median metallicity after corrections for dust extinction, photometric bias, and completeness. The shaded black regions indicate the crowded central region where our bias and completeness corrections fail in recovering the true metallicity distribution (Section~\ref{sec:completeness-limit}). Included as a reference point, the black ellipse is a ring with with constant deprojected radius at 10~kpc. Our measurement of the metallicity gradient uses these corrected median metallicities (Section~\ref{sec:metallicity-gradient}). {\em Middle Panel}~-- on top of the median metallicity map, we plot green contours at constant RGB density as inferred from the IR photometry (\cite{Dalcanton2012}, Seth~et~al.~{\em~in~prep}). Additionally, a solid black line shows Andromeda's bar length and azimuthal offset measured by \cite{Athanassoula2006}. We observed a region of high median metallicity correlated with the RGB density and bar. In Section~\ref{sec:metal-rich-bar}, we argue these high metallicities are likely a result of bar interaction. {\em Bottom Panel}~-- we overplotted the median metallicity map with the average reddening (\avfred{}). Every spatial bin below the lowest dust contour we define as low extinction (\avfred{}$ < 0.25$~mag). We base our results on these low extinction areas. This figure shows the high metallicity region we associate with Andromeda's bar is not caused by dust. }

\caption{ The median metallicity gradient in M31. To create this figure, the median metallicity map (Figure~\ref{fig:corrected-median-metallicity-map}) was divided into annular bins with $b/a=0.275$; within these bins the mean of the map is shown as the solid line, while the shaded region shows the standard deviation of the pixels in the bin. High extinction regions (\avfred{} $> 0.25$) were excluded. The red line shows the gradient including corrections for dust extinction and photometric bias and completeness, while in cyan the metallicities have only been corrected for dust extinction. The bias and completeness correction changes the gradient by $\sim$1$\sigma$, while the dust extinction correction changes the gradient by less than a 0.001$\sigma$. The black line and shaded region shows the best linear fit at deprojected radii $R > 4.45$~kpc~($\frac{1}{2} R_{eff}$; \cite{Courteau2011}) to the corrected metallicities, the best-fit slope is -0.020$\pm$0.004 dex/kpc. The black shaded region at $R \lesssim 4$~kpc where our RGB data becomes incomplete at high metallicities. }

253 — 1511.00304

\caption{ (Color online) Mass distributions in cell 3 at three different temperatures. Black lines: NSE calculation with pairing. Red lines: perturbative inclusion of the in-medium energy shift. {\cred If we want to discuss in-medium effects, these figures will have to be improved.} }

254 — 1511.00463

\caption{\label{fig:Shower2}Illustration of the intrinsic variability of shower development. {\bf Top:} Simulation of 10 showers, each initiated by a \gray\of$300\U{GeV}$. {\bf Bottom:} Simulation of 10 showers initiated by a proton of the same energy. Due to larger transverse momentum transfers, hadronic showers show larger fluctuations. From~\cite{TheseMathieu}.}

255 — 1511.00923

\caption{Characteristic Brownian diffusion time and the parameters $A(q)$, $B(q)$ versus $q$. (a) $\tau_\text{d}$ ($\circ$) and its fit (\textcolor{red}{\textbf{---}}) using Eq.\ref{eq:relax}. (b) ($A$ (\textcolor{blue}{+}), $B$ (\textcolor{blue}{.})) measured with strategy 1 and ($A$ ($\Box$), $B$ ($\diamond$)) fitted with strategy 2. The red dashed line is a fit of $A(q)$ following the model in Ref\citep{3_DDM3D}. The vertical black dash lines delimite regime 1, 2 and 3. Fits are performed in regime 2.}

\caption{Fit parameters for the motile bacteria as a function of $q$ as extracted from the fit of the measurements of the DDM matrix $\mathcal{D}$. (a) characteristic time for the diffusion, $\tau_d$ (\textcolor{red}{$\circ$}), and balistic motion, $\tau_r$ (\textcolor{blue}{$\circ$}). (b) Fraction of motile bacteria $\alpha$. (c) $Z$. Inset: Shulz velocity distribution, $P$ for $<Z>$=2.4. (d) Standard deviation of the Shulz distribution $\sigma$ obtained from Eq. \eqref{eq:sig}.}

256 — 1511.00962

\caption{ {\color{red}Spectral function} on the impurity site, for different numbers of sites in two identical leads. Left: The exact $A$ from the finite system. Right: $A$ using the implantation method. }

257 — 1511.01493

\caption{ %\textcolor[named]{Blue}{Triplicate redundancy} of each diggram Emergence of degeneracy. \textcolor[named]{Blue}{$C$}, say, denotes \textcolor[named]{Blue}{$\mathcal{M}_C - \mathcal{S}_C$}.}

\caption{An identical diagram of $\left(2_E,\,2\right)$-type is generated in three ways from $\textcolor[named]{Blue}{\mathcal{M}_C}$ (left) and $\textcolor[named]{Magenta}{\mathcal{M}_D}$ (middle,\,right). % \textcolor[named]{Red}{The red stuffs are generated by \textcolor[named]{Red} {The red propagators and vertices are generated by the ensemble average of (${\rm QCD} + {\rm QED}$) }. } \label{fig:(2E,2)} \end{center} \vspace*{-\intextsep} \end{figure} %%%%% The term $\left(\textcolor[named]{Red}{-\mathcal{K}_{D}}\right)$ in Eq.~(\ref{eq:nonperturbativeQEDMethod}) is the one added here to subtract the unwanted $O(\alpha^3)$ HVP contribution contained in the other terms. To construct $\textcolor[named]{Red}{\mathcal{K}_{D}}$, we prepare two sets of $\left({\rm QCD},\,{\rm QED}\right)$. Practically, they may be two independent important samples of a pair of $\left(U,\,A\right)$ generated by dynamical (${\rm QCD} + {\rm QED}$) simulation. The quark in the upper loop on the right-hand side of Fig.~\ref{fig:D+KD} is charged only with respect to the first $\left({\rm QCD},\,{\rm QED}\right)$ and the one in the lower loop only with respect to the second $\left({\rm QCD}

258 — 1511.01658

\caption{Schematic illustration of steady state constrained optimisation problem. The path of a continuous analogue of a constrained optimiser ({\color{darkgreen}\bf---}) from a starting point $(\theta_0,x_{s,0})$ to a local minimum $(\theta^*,x_s(\theta^*))$ on the steady state manifold ({\color{darkred}\bf---}) is depicted.}

259 — 1511.02013

\caption{\label{fig:Si10pop}Calculated fractional $^{29}$Si$^{10+}$ level populations as a function of storage time using theoretical transition rates \cite{Wang2015} (see also \tref{tab:Si10levels}) for the 116 lowest Si$^{10+}$ levels. In addition, the rate for the hyperfine induced $2s\,2p\;^3P_0 \to 2s^2\;^1S_0$ transition ($A_\mathrm{HFI} = 0.06011$~s$^{-1}$) was taken from \cite{Cheng2008a}. The different lines correspond to the following levels: $2s^2\;^1S_0$ (\full), $2s\,2p\;^3P_0$ (\broken), $2s\,2p\;^3P_1$ (\dashddot), and $2s\,2p\;^3P_2$ (\chain). The dotted curve (\dotted) is the sum of the fractional populations of the remaining 112 levels. }

260 — 1511.02228

\caption{We largely improve ({\color{red} red}) over the original example-based single image super-resolution methods ({\color{blue} blue}), \ie our IA method is 0.9dB better than A+~\cite{Timofte-ACCV-2014} and 2dB better than Yang~\etal~\cite{Yang-CVPR-2008}. Results reported on Set5, $\times3$. Details in Section~\ref{ssc:generality}.}

\caption{Average PSNR on Set5, Set14, and B100 and the \textcolor{red}{improvement (red)} of \textcolor{blue}{our IA (blue)} over \textbf{A+ (bold)} method. }

261 — 1511.02317

\caption{\label{fig:Nonlinear-z}(Color online) (a) Nonlinear Zeno parameter with $\frac{\Gamma_{a}}{k}=\frac{\Gamma_{b}}{k}=10^{-2},\,\frac{\Delta k_{a}}{k}=1.1\times10^{-3},\,\frac{\Delta k_{b}}{k}=1.1\times10^{-3}$ with $\alpha=5,\,\beta=3,$ and $\gamma=2,\,\delta=1$ and -1 in stimulated case (smooth and dashed lines) and $\gamma=-2,\,\delta=1$ and -1 in dot-dashed and dotted lines. (b) $\alpha=10,\,\gamma=3,\,\delta=1$ with $\beta=3,6,$ and 7 for smooth, dashed and dot-dashed lines, respectively. (c) shows the change in nonlinear Zeno parameter with changing the phase of $\alpha$ or $\beta$ by an amount of $\pi$\textcolor{red}{{} }for $\alpha=\beta=6,$ and $\gamma=3,\,\delta=2$. It is also observed that the change in phase of $\alpha$ is equivalent to change in phase of $\beta$.}

262 — 1511.02386

\caption{% \textbf{(a)} $q(\mblambda;\mbtheta)$: The (reparameterized) natural parameters assume a multivariate prior, with different areas indicated by \red{red} and \blue{blue}. \textbf{(b)} $\prod_{i=1}^2 q(z_i\g\mblambda_i)$: The latent variables are drawn from a mean-field family, colorized according to its drawn parameters' color. \vspace{-2ex} }

263 — 1511.03257

\caption{Mean Average Precision and training time @ 32 bits for Cifar-10. \textbf{Bold} denotes the best performing method while {\color{red}{red}} denotes the best online technique.}

\caption{Mean Average Precision and training time @ 32 bits for Sun397. \textbf{Bold} denotes the best performing method while {\color{red}{red}} denotes the best online technique.}

264 — 1511.03611

\caption{\blue Table of Notation }

265 — 1511.03814

\caption{\label{fig:Distribution-of-baseline}Distribution of baseline performance for various action classes. We note the typical size of the action-objects for the transitive actions (those involving objects), and classify them as small(\textcolor{red}{red}), large(\textcolor{green}{green}), and leave the rest of the classes undecided (orange). Mean performance is significantly worse on actions involving small objects (37.5 \%) vs. large objects (87 \%).}

\caption{\label{fig:Flow-of-proposed}Flow of proposed method. A fully convolutional network is applied to predict body parts and action objects. This guides a secondary network to refine the predictions on a few selected image regions. Using contextual features, the regions are ranked \& pruned. The remaining candidates are scored using both global and local features to produce a final classification, along with a visualization of the detected action object.}

266 — 1511.04048

\caption{Visualization of the \emph{direction} of net force and object velocity. The \textcolor{green}{velocity} is shown in green and the \textcolor{magenta}{net force} is shown in magenta. The corresponding Newtonian scenario is shown above each image. }

267 — 1511.04128

\caption{\label{FigNew1} Simulated signals of length 512 ({\color{mygray}----}), from the widely linear complex autoregressive process~\eqref{eq:ICAR} in (a) and (c), and from an order one improper process~\eqref{eq:doublywhite} in (b) and (d). In (a) and (b), the parameters are $\lambda=0.99$, $\alpha=\pi/6$, $\sigma^2_\nu=1$, and $c_\nu=\exp(3\pi\mathrm{i}/4)/5$. In (c) and (d), the parameters are $\lambda=0.297$, $\alpha=\pi/6$, $\sigma^2_\nu=1$, and $c_\nu=2\exp(3\pi\mathrm{i}/4)/3$. For the widely linear complex autoregressive process~\eqref{eq:ICAR} in (a) and (c), we additionally set $\gamma=0.099$ and $\phi=-\pi/4$, which are set to zero in (b) and (d) for the model~\eqref{eq:doublywhite}. Expected second moment ellipses for the standard deviation are also given ({- - -}), which are calculated from Section~\ref{S:autocov}.}

\caption{\label{Fig2}Panel (a) displays the periodogram ({---}) and model fit ({\color{red}- - -}) of the signal $Z_t$ displayed in Fig.~\ref{Fig1}(c). Panel (b) shows $Z_t$ in the interval (UTC) 16:33:46 to 16:36:26 ({---}) and 16:38:19 to 16:40:59 ({\color{mygray}---}). The periodogram ({---}) and model fit ({\color{red}- - -}) are displayed for $Z_t$ in the intervals (c) 16:33:46 to 16:36:26 and (d) 16:38:19 to 16:40:59. In (a), (c), and (d), parameter estimation is performed for $\omega\in[-\pi/4,\pi/4]$, as indicated by the vertical dashed boundaries, and we extend the fitted lines to all frequencies ({\color{blue}-$\cdot$-}).}

\caption{\label{Fig4}Estimates of (a) the eccentricity, $\varepsilon$, and (b) the orientation, $\psi$, in radians, of the seismic signal of Fig.~\ref{Fig1}. (----) are the estimates from fitting a widely linear complex autoregressive process of order one across a 161-second sliding window, with the 95$\%$ confidence intervals given in gray. (\textcolor{blue}{-~$\cdot$~-}) are the estimates from the method of~\cite{lilly2010bivariate}, and (\textcolor{red}{- - -}) are the estimates obtained from the Fourier Transform evaluated at the two frequency peaks. All estimates and confidence intervals, including the nonparametric techniques, have been smoothed in time using a moving average window of width 11.}

\caption{\label{Fig5}The likelihood ratio statistic $W$ ({\color{mygray}----}) over time, smoothed with a moving average window of width 11, from fitting the proper~\eqref{eq:CAR} and widely linear improper~\eqref{eq:ICAR} complex autoregressive order one processes to $Z_t$ across a 161-second sliding window. The signal is then divided into 11 non-overlapping windows, as indicated by the vertical gray-dashed lines, where the vertical black solid lines indicate the analysis window of Figs. 4--6. The likelihood ratio statistic for each of these windows is then given by (----). We also display the 95th percentile of a $\chi^2_2$ distribution (- - -), and after applying a False Discover Rate (FDR) procedure to control the rate of false positives, we reject propriety in each segment except the first two.}

268 — 1511.04176

\caption{TSNE plots for feature representation of unique words starting with top eight most frequent alphabets({\color{blue}S}, {\color{darkgreen}T}, {\color{red}W}, {\color{cyan}H}, {\color{magenta}B}, {\color{yellow}C}, {F}, {\color{brown(web)}A} including lowercase. (Color based on first alphabet of words)}

\caption{TSNE plots for feature representation of unique words starting with S and top eight \textit{second} alphabets ({\color{blue}t}, {\color{darkgreen}h}, {\color{red}o}, {\color{cyan}e}, {\color{magenta}i}, {\color{yellow}u}, {a}, {\color{brown(web)}p}) in word with respect to population in S. (Color based on second alphabet of words starting with S.)}

269 — 1511.04203

\caption{Runtime breakdown of a single Picard equilibrium-iteration step (average over 4 iteration steps times 1000 time points) for different combinations (subset of Table~\ref{tab:pide_timing}) of numerical resolution and computational resources. The individual algorithmic steps from top to bottom correspond to the sequence in Sect.~\ref{sect:algorithm_equil}. Routines labelled with asterisks are dominated by MPI communication ($^*${\tt MPI\_Allgather}, $^{**}${\tt MPI\_Allreduce}).}

270 — 1511.04491

\caption{Benchmark results. Average PSNR/SSIMs for scale factor $\times$2, $\times$3 and $\times$4 on datasets Set5, Set14, B100 and Urban100. {\color{red}Red color} indicates the best performance and {\color{blue}blue color} refers the second best.}

271 — 1511.04587

\caption{Scale Factor Experiment. Several models are trained with different scale sets. Quantitative evaluation (PSNR) on dataset `Set5' is provided for scale factors 2,3 and 4. {\color{red}Red color} indicates that test scale is included during training. Models trained with multiple scales perform well on the trained scales. }

\caption{Average PSNR/SSIM for scale factor $\times$2, $\times$3 and $\times$4 on datasets Set5, Set14, B100 and Urban100. {\color{red}Red color} indicates the best performance and {\color{blue}blue color} indicates the second best performance.}

272 — 1511.04674

\caption{Sample classified with important words highlighted. Words in red affect the price of the classified negatively (e.g. \textcolor[rgb]{1.0,0.0,0.0}{Jumeirah}), while words in blue affect the price positively (e.g. \textcolor[rgb]{0.0,0.0,0.6235294117647059}{Palm}).}

273 — 1511.04862

\caption{Probability density function of $(a)~\theta_W$ and $(b)~\theta_{77}$ measured at the top (\textcolor{blue}{$\circ$}), middle (\textcolor{black}{$\lozenge$}) and bottom (\textcolor{red}{$\vartriangle$}) heights of the respective LSCs.}

274 — 1511.05045

\caption{ Schematic description of IDT as procedures of multiple convolution and pooling operations. Dashed \textcolor{red}{red} and \textcolor{green}{green} boxes represent the procedure of generating handcrafted descriptors and KMeans-based BoW encoding, respectively. In each operation, the first three numbers are the receptive field sizes in space and time (x, y, t) and the last number indicates the size of output channels. }

\caption{ The neural network architecture of an ISA network with PCA preprocessing. The dashed \textcolor{blue}{blue} boxes represent the outputs of our model.}

275 — 1511.05517

\caption{\footnotesize \label{fig2} (Color online) (a) Typical optical images in $xy$-plane with a rod confined by a field. Overlaid are the centroid ({\color{magenta}$\bullet$}), length and orientation ({\color{cyan}--}) and the ferromagnetic cap ({\color{red}$\bullet$})of the rod. (b) A short segment of projected length $L_p(t)$ demonstrates the hopping behaviour.(c) Scattered [$\theta(t)$,$\phi(t)$] for no field ({\color{red}$\circ$}) and $140$ G field ({\color{NavyBlue}$\bullet$}) demonstrates the magnetic trapping (for clarity, we display every tenth sequential point). (d) Histogram of $\phi(t)$ for $B=140$ G case enables us to calculate the magnetic moment's offset angle $\epsilon$ by measuring the shift of the peak from $\pm \pi/2$.}

\caption{\footnotesize (Color Online) (a) Histogram of $\theta(t)$ shows a markedly increased tendency for vertical states (small $\theta$) to be realised when a magnetic field is present ({\color{NavyBlue}$\bullet$}) with respect to a free rod ({\color{red}$\bigtriangleup$}). The curves represent absolute probability weightings $P(\theta) \sin(\theta)$ where we have numerically integrated Eq.\(\ref{corrected}) using independently measured parameters $\mu$, $\epsilon$, $m$, $L$ and $d$. (b) The relative distribution $P(\theta | 140 $G$)/P(\theta | 0 $G$)$ agrees well with the theoretical calculations, showing an $O(10)$ increase in likelihood for small-$\theta$ states due to the magnetic field compared to no field.}

276 — 1511.05549

\caption{\label{Tab_ion_energies} Standard chemical potential ($\mu^0$, in eV) of aqueous ions at 298.15 K obtained from four experimental databases. The Pourbaix \& Bard's data (column 1) are used in this work. Burgess's and Lide's data are calculated from the redox potentials listed in the provided references.}

277 — 1511.05596

\caption{ \blue{ Ring model: Symmetric samples restricted internally and externally by equilateral triangles. We keep the external radii constant and equal $R_{\mathrm{ext}}=50$ nm and change the side thickness $d$. In Fig. (a) $d=10$ nm and in Fig. (b) $d=20$ nm.} The areas included between the internal and external boundaries where effective quantum wells are formed are marked in red or turquoise. Doted turquoise and red lines indicate internal boundaries and corner areas of $20$ (a) and $10$ nm (b) thick samples. }

\caption{\blue{Probability distribution associated with the ground state (a-d), the second energy level (e-h) and the third level (i-l) for symmetric triangular rings of different side width $d$. The $x$ and $y$ coordinates are in units of $R_{\mathrm {ext}}$ and probability distribution in units of $R_{\mathrm {ext}}^{-2}$. The distributions for d=8 nm are taken as reference, and for the other cases they are magnified with factors shown on the top left corner of each graph.} }

\caption{ \blue{Energy levels of non-symmetric triangular rings.} % the samples were % restricted externally by the same % equilateral triangles for which $R_{\mathrm{ext}}=50$ nm % while the internal boundaries % were defined by different side thicknesses. The values of the % side thickness shown in the figures refer to the side parallel % to lines $y=(-x+1)/\sqrt{3}$, the thicknesses of the two other % sides were increased by $4\%$ (sides parallel to the $y$ axis) and % $8\%$ (sides parallel to line $y=(x-1)/\sqrt{3}$) in Fig.\ \ref{fig:sample} % with respect to the values shown in figures (a) and on % the axises of Figs. (b) and (c). (a) The six lowest energy levels and their degeneracy for two samples with aspect ratios equal to $0.16$ \blue{($d=8$ nm)} and $0.36$ \blue{($d=18$ nm)}, respectively. (b) Ground state and the fourth level changing with side thickness. (c) Energy gap dependence on side thickness: energy splitting between the two lowest levels associated with corner localized probability distributions, $\Delta_{21}$, an energy gap separating corner- and side-localized states, $\Delta_{43}$, and the energy difference between the fourth (mostly side-localized \blue{level}) and the ground state, $\Delta_{41}$. \red{ The values of the side thickness shown in Fig. (a) and on the axises of Figs. (b) and (c) refer to the thickness of the narrowest side.} }

\caption{Probability distribution associated with the ground state (a-d), the third energy level (e-h) and the fourth one (i-l) for non-symmetric samples. \red{ The samples were restricted externally by the same equilateral triangles for which $R_{\mathrm{ext}}=50$ nm while the internal boundaries were defined by different side thicknesses. The values of the side thickness shown in the figure captions refer to the sides parallel to lines $y=(-x+1)/\sqrt{3}$, the thicknesses of the two other sides were increased by $4\%$ (sides parallel to the $y$ axis) and $8\%$ (sides parallel to line $y=(x-1)/\sqrt{3}$) with respect to the shown values.} \blue{The $x$ and $y$ coordinates are in units of $R_{\mathrm {ext}}$ and probability distribution in units of $R_{\mathrm {ext}}^{-2}$. The distributions for $d=8$ nm are taken as reference, and for the other cases they are magnified with factors shown on the top left corner of each graph.} }

\caption{Absorption coefficients associated with clockwise (green solid) and counterclockwise (red dashed) polarization superimposed on the density of states \blue{(gray dotted)} for a 12 nm thick equilateral ring \blue{obtained} in the presence of external magnetic field. Values of the field shown in the right panels refer also to the left ones. Figs. (a), (c) and (e) show transitions from the ground state to other corner-localized states while Figs. (b), (d) and (f) transitions to side-localized states. For visibility we use a logarithmic scale for the absorption functions. }

\caption{Absorption coefficients associated with clockwise polarization and excitations from the ground state to other corner- (a) and side-localized (b) states of equilateral triangular rings \blue{of different side thickness $d$. Values shown in Fig. (b) refer to both panels.} $B=0.53$ T. For visibility we use a logarithmic scale for the absorption functions. }

\caption{Absorption coefficients associated with clockwise polarization and excitations from the ground state to the first five excited \blue{levels} of non-symmetric samples, defined in Figs.\\ref{fig:energies_non} and \\ref{fig:localization_non}. The numbers shown in the figure refer to the thickness of the narrowest sides, $B=0$. For visibility we use a logarithmic scale for the absorption functions. }

278 — 1511.05966

\caption{\label{fig2} Mean vertical temperature profiles, $\oT(z)$, in the static state (\dashedrule) and in 3D simulations with $Pr=1$ and $R=10^6,10^7,10^8,10^9,10^{10}$ (from right to left).}

\caption{\label{fig:IQs} Variation of mean quantities with $R$ and $Pr$ in 2D (\textcolor[rgb]{1,0,0}{\tiny{$\blacksquare$}}) and 3D (\textcolor[rgb]{0,0,1}{\large{$\bullet$}}) simulations. The fraction of internally produced heat flowing outward across the bottom boundary, $\fB$, is shown (a) for various $R$ with $Pr=1$ and (b) for various $Pr$ with $R=2\e7$. For the same simulations, the dimensionless mean fluid temperature, $\T$, is shown (c) for various $R$ (compensated by $R^{-1/5}$) and (d) for various $Pr$. The data shown are tabulated in the Appendix, except for the 2D data in panels $a$ and $c$, which are from \citet{Goluskin2012}.}

\caption{\label{fig:KE} Mean vertical profiles of root mean square horizontal velocity (top) and vertical velocity (bottom) for 2D ({\color{red}$\dashedrule$}) and 3D ({\color{blue}$\solidrule$}) simulations with $Pr=1$ and $R=10^6,10^8,10^{10}$. In the 2D cases, $v\equiv0$. The 3D profiles are qualitatively consistent with those reported by \citet{Worner1997}.}

\caption{\label{fig:N} Variation of $\wt{N}$ (a) with $\wt{Ra}$ (compensated by $\wt{Ra}^{1/4}$) and (b) with $Pr$ in 2D (\textcolor[rgb]{1,0,0}{\tiny{$\blacksquare$}}) and 3D (\textcolor[rgb]{0,0,1}{\large{$\bullet$}}) simulations. The $\wt{N}$ and $\wt{Ra}$ values are calculated according to \eqref{eq: Nwt} from the $\T$ values represented in figures \ref{fig:IQs}c and \ref{fig:IQs}d.}

279 — 1511.06644

\caption{The generated motion with a step function signal, starting with walking (\textcolor{blue}{blue}), switching to running (\textcolor{red}{red}) and switching back to walking (\textcolor{blue}{blue}).}

280 — 1511.06783

\caption{\textcolor{red}{新しい物に差し替え}}

281 — 1511.06789

\caption{ \textbf{Left:} Classifier confidence versus false positive rate on 100,000 randomly sampled from Flickr images (YFCC100M~\cite{thomee2015yfcc100m}) with dog detections. Even the most confident images have a 20\% false positive rate. \textbf{Right:} Samples from Flickr. Rectangles below images denote correct ({\color[rgb]{0,0.8,0} green}), incorrect ({\color{red} red}), or ambiguous ({\color[rgb]{0.8,0.8,0.0}yellow}). \textbf{Top row:} Samples with high confidence for class ``Pug'' from YFCC100M. \textbf{Bottom row:} Samples with low confidence score for class ``Pug''. }

282 — 1511.06838

\caption{ User generated ANP tags of images are inherently noisy: The same noun ({\it \textcolor{blue}{baby}}) could mean different entities, and a positive adjective ({\it \textcolor{red}{attractive}}) could modify the pairing noun with a negative sentiment when used sarcastically. }

283 — 1511.06864

\caption{(Color online) Deviation $\Delta$, Eq.~(\ref{eq:delta}), vs the truncation order $n$, for the spin chain model, Eqs.~(\ref{eq:spin1}-\ref{eq:spin3}). Three values of the driving frequency, $\hbar\Omega=4.2h_z$ ({\color{red}$\bullet$}), $4.6h_z$ ({\color{blue}$\subHalmos$}), and $5.0h_z$ ($\ast$) were used in simulations, while keeping the ratio $\xi/\hbar\Omega=2/3$ fixed. Arrows indicate the effective truncation order $n_{\rm{eff}} = 13$ ($\hbar\Omega = 4.2h_z$), $14$ ($\hbar\Omega = 4.6h_z$) and $16$ ($\hbar\Omega = 5.0h_z$).}

\caption{(Color online) The dependence $\Delta{\rm Prob}\equiv \Delta{\rm Prob}^{[0]}$, Eq.~(\ref{difference}), on the system-bath coupling (squared) $\lambda^2$ (dissipation strength) for three values of the driving frequency, $\hbar\Omega=4.2h_z$ ({\color{red}$\bullet$}), $4.6h_z$ ({\color{blue}$\subHalmos$}), and $5.0h_z$ ($\ast$). There is a resonance (see text) for $\hbar\Omega=4.6h_z$, which is responsible for a strong deviation from the effective Floquet-Gibbs state in the weak-coupling limit. The deviation decreases upon the increase of the coupling strength. The cutoff frequency, Eq.~(\ref{eq:cutoff}), $\hbar\omega_c = 100h_z$. Other parameters are same as in Fig.~\ref{sx}.}

\caption{(Color online)The populations ${\rm P}_k$ of the eigenstates of the effective Floquet Hamiltonian of the driving frequency $\hbar\Omega=4.6h_z$ and two values of the coupling, $\lambda^{2} = 10^{-6}$ ({\color{red}$\bullet$}) and $\lambda^2=10^{-2}$ ($\circ$). In case of $\lambda^{2}=10^{-6}$ there are two eigenstates resonantly coupled by the driving field. Their populations deviate strongly from the Boltzmann distribution. Solid black line corresponds to the Boltzmann distribution with $\beta_{\rm eff}=0.946h_z^{-1} < \beta = 1h_z^{-1}$. Other parameters are same as in Fig.~\ref{DIFGa}.}

\caption{(Color online) The difference $\Delta{\rm Prob}\equiv \Delta{\rm Prob}^{[0]}$, Eq.~(\ref{difference}), as a function of the cutoff frequency $\omega_c$, of the spectral density of the heat bath, Eq.~(\ref{eq:cutoff}), for three values of the driving frequency, $\hbar\Omega=4.2h_z$ ({\color{red}$\bullet$}), $4.6h_z$ ({\color{blue}$\subHalmos$}), and $5.0h_z$ ($\ast$) at $\lambda^2=10^{-2}$. Gray stripe marks the interval $\hbar\omega_c/h_z \in [4.2, 5]$.}

\caption{(Color online) The dependence of $\Delta{\rm Prob}\equiv \Delta{\rm Prob}^{[0]}$, Eq.~(\ref{difference}), on the system-bath coupling $\lambda^2$ for $\hbar\Omega=4.2h_z$ ({\color{red}$\bullet$}), $4.6h_z$ ({\color{blue}$\subHalmos$}), and $5.0h_z$ ($\ast$). The cutoff frequency is $\hbar\omega_c = 0.4h_z$. For $\hbar\Omega=4.6h_z$ there is a resonant coupling of two states induced by the driving field, so that the dependence follows the scenario presented on Fig.~\ref{DIFGa}. Inset: The populations ${\rm P}_k$ of eigenstates of the effective Hamiltonian for $\hbar\Omega=4.6h_z$, for two values of dissipation strength, $\lambda^2=10^{-2}$({\color{blue}$\subHalmos$}, blue closed boxes) and $\lambda^2=10^{-6}$({\color{green}\protect\makebox[6pt]{$\square$}{\hspace{0.em}}}, green open boxes). The solid line is the Boltzmann distribution with the exponent $\beta_{\rm eff} = 0.99$. %The values of $n_{{\rm eff}}$ is same as that in Fig.~\ref{sx}. }

\caption{Probability difference $\Delta {\rm Prob} \equiv \Delta {\rm Prob}^{[0]}$ vs the truncation frequency $\omega_{\rm trunc}$ at $\hbar \Omega =4.6 h_z$ for sets of system-bath coupling $\lambda^2$ and cutoff frequency $\omega_c$, $(\lambda^2, \hbar\omega_c)=(10^{-2}, 100 h_z)$ ({\color{red}$\bullet$}), $(\lambda^2, \hbar\omega_c)=(10^{-6}, 100 h_z)$ ({\color{blue}$\subHalmos$}), $(\lambda^2, \hbar\omega_c)=(10^{-2}, 0.4 h_z)$ ($\ast$), and $(\lambda^2, \hbar\omega_c)=(10^{-6}, 0.4 h_z)$ ({\color{green}$\blacktriangle$}). }

\caption{ Time dependence of probability differences $\Delta{\rm Prob}^{[t]}$ for $\hbar\Omega=4.6 h_z$. Left figure shows data for $\lambda^2=10^{-6}$ ({\color{red}$\bullet$}), $\lambda^2=2^6*10^{-6}$ ({\color{blue}$\subHalmos$}), and $\lambda^2=2^{14}*10^{-6}$ ($\ast$) at $\hbar \omega_c=100 h_z$. Right figure shows data for $\hbar \omega_c=0.4 h_z$ ({\color{red}$\bullet$}), $\hbar \omega_c=3.2 h_z$ ({\color{blue}$\subHalmos$}), and $\hbar \omega_c=100 h_z$ ($\ast$) at $\lambda^2=0.01$. Thus, the probability differences are almost independent of $t$. }

284 — 1511.06920

\caption{{\bf Cell activity characterization.} \red{a) Firing rates $\nu_i$ of 6 selected neurons belonging to two anti-correlated assemblies, the color identifies the assembly and the colors correspond to the one used in b) for the different clusters;} b) raster plot activity, the firing times are colored according to the assembly the neurons belong to; c) cross-correlation matrix $C(\nu_i,\nu_j)$ of the firing rates. \red{The neurons in panel b) and c) are clustered according to the correlation of their firing rates by employing the \textit{k-means} algorithm; the clusters are ordered in terms of their average correlation (inside each cluster) from the highest to the lowest one (for more details see Methods).} The firing rates are calculated over overlapping time windows of duration 1 s, the origins of successive windows are shifted by 50 ms. The system is evolved during $10^7$ spikes, after discarding an initial transient of $10^5$ spike events. Other parameters used in the simulation: $g = 8$, $K = 20$, $N = 400$, $k_{mean} = N_{act}/15$, $\Delta V = 5$ mV and $\tau_{\alpha} = 20$ ms. \red{The number of active neurons is 370, corresponding to $n^* \simeq 93$ \%.}}

\caption{\textbf{Metrics of structured activity vs lateral inhibition strength.} \red{a) Metrics entering in the definition of $Q_0$ versus the synaptic strength $g$. From top to bottom: Averaged coefficient of variation $\langle CV \rangle_N$, standard deviation of the cross-correlation matrix $\sigma(C)$,} and the fraction of active neurons $n^*$. The solid (dashed) line refers to the case $\Delta V = 1$ mV ($\Delta V= 5$ mV). The minimum number of active neurons is achieved at $g=g_{min}$, this corresponds to a peak amplitude of the PSP $A_{PSP} = 0.064$ mV ($A_{PSP} = 0.184$ mV) for $\Delta V = 1$ mV ($\Delta V = 5$ mV) (for more details see Methods). b) Distributions $P(\overline{ISI})$ of the average ISI for a fixed $\Delta V= 5$ mV and for two different coupling strengths, $g=4$ (red triangle-up symbol) and $g=10$ (blue triangle-down). Inset, the distribution $P(CV)$ of the $CV$ of the single neurons for the same two cases. c) $Q_0$ and $Q_d$, as defined in Eqs. \eqref{eq:Q0Indicator} and \eqref{eq:deltaM2}, versus $g$ for $\Delta V = 1$ mV. d) Same as c) for $\Delta V= 5$ mV. \red{Other parameters as in Fig. \ref{fig:PonziBenchmark}}}

\caption{\textbf{Metrics of structured activity vs post-synaptic time duration.} a) Metrics $Q_0$ (in solid line) and $Q_d$ (dashed) as a function of \red{the pulse time scale for the parameter values $\{\Delta V,g\} = \{5 \enskip {\rm mV},8\}$ corresponding to the maximum $Q_0$ value in Fig. \ref{fig:Q0andCV}(d)}. Probability distribution functions $P(CV)$ ($P(CV_2$)) for the coefficient of variation $CV$ (local coefficient of variation $CV_2$) are shown in b) (in c)) for three representative $\tau_{\alpha} = \{2, 9, 20\}$ ms, displayed by employing the same symbols and colors as indicated in a). \red{For these three cases the average firing rate in the network is $\langle \nu \rangle = \{8.81, 7.65, 7.35\}$ Hz ordered for increasing $\tau_{\alpha}$-values}. Other parameters as in Fig. \ref{fig:PonziBenchmark}}

\caption{\textbf{Single neuron statistics.} \red{First row : distributions $P(ISI)$ for one representative cell in the network are shown in black. Second row: the corresponding Poissonian reconstruction of the $P(ISI)$ are reported in red. In all plots the main figure displays the distributions at short ISIs, while the inset is a zoom out of the whole distribution. Third row: single neuron distribution of the $CV^{(i)}_2$ for the considered neuron (black solid lines with circles) and its Poissonian distribution (red dashed line with squares). The left (right) column corresponds to $\tau_\alpha = 20$ (2 ms).} The network parameters are $\Delta V = 5$ mV and $g = 8$, and the others as in Fig.~\ref{fig:PonziBenchmark}, both the examined neurons have $I_s = -45.64$ mV. For the Poissonian reconstruction the frequencies of the incoming uncorrelated spike trains are set to \red{$\langle \nu \rangle_N \approx 7.4$ Hz ($\langle \nu \rangle_N \approx 8.3$ Hz)} for $\tau_{\alpha} = 20$ms ($\tau_{\alpha} = 2$ms), as measured from the corresponding network dynamics. The distributions are obtained by considering a sequence of $10^9$ spikes in the original network, and $10^7$ events for the Poissonian reconstruction. }

\caption{\red{\textbf{Cell assemblies and connectivity.} a) Cross-correlation matrix $C(\nu_i,\nu_j)$ of the firing rates organized according to the clusters generated via the \textit{k-means} algorithm with $k=15$, the clusters are ordered as in Fig.~\ref{fig:PonziBenchmark}(c) from the highest to the lowest correlated one. b) Connectivity matrix $\mathcal{C}_{ij}$ with the indices ordered as in panel a). Here, a black (copper) dot denotes a 1 (0) in $\mathcal{C}_{ij}$, i.e. the presence of a synaptic connection from $j$ to $i$. c) Average cross-correlation $\langle C \rangle_{ml}$ among the elements of the matrix block $(m,l)$ , versus the probability $p_{ml}$ to have synaptic connections from neurons in the cluster $l$ to neurons in the cluster $m$. Black squares indicate the blocks along the diagonal delimited by black borders in panel a) and b) , i.e. they correspond to the pairs $\{ \langle C \rangle_{mm}, p_{mm} \}$; blue triangles denote the ten blocks with the lowest $\langle C \rangle_{ml}$ values, which are also delimited by blue edges in a) and b). The vertical red dashed line indicates the average probability to have a connection $p = 5 \%$. The black solid line is the linear regression to the data ($ \langle C \rangle_{ml} \approx 0.15 - 3.02 p_{ml}$, correlation coefficient $R = -0.72$). Other parameters as in Fig. \ref{fig:PonziBenchmark}.}}

\caption{\textbf{Sequential switching.} a) Raster plot associated to the two input protocols $I^{(1)}$ and $I^{(2)}$. The circles denote the clusters of active neurons appearing repetitively after the presentation of the stimulus $I^{(1)}$. Vertical lines denote the switching times between stimuli. b) Averaged State Transition Matrix $\overline{D}$ , \red{obtained by considering a $4 T_{sw} \times 4 T_{sw}$ sub-matrix averaged over $r=5$ subsequent time windows of duration $4 T_{sw}$} (see the section \textit{Methods} for details). The inputs $I^{(1)}$ and $I^{(2)}$ are different realization of the same random process, they are obtained by selecting $N$ current values $I_i$ from the flat interval $[V_{th}, \, V_{th}+\Delta V]$. The input stimuli are switched every $T_{sw} = 2$ s. Number of clusters $k = 35$ in a). Other parameters as in Fig. \ref{fig:PonziBenchmark}.}

\caption{{\bf Pattern separation.} \red{Average dissimilarity as a function of the fraction $f$ of inputs differing from the control input,} for the values of $\tau_{\alpha} = 20$ms (black circles) and $\tau_{\alpha} = 2$ms (red squares) with two different observation windows $T_{E} = 2$s (solid line) and $T_{E} = 10$s (dashed line). Other parameters used: $\Delta T= 50$ms, $\Delta V= 5$ mV. Remaining parameters as in Fig. \ref{fig:PonziBenchmark}.}

\caption{\textbf{Computational capability of the network.} Characterization of the firing activity of the network, obtained as response to three consecutive inputs presented in succession. a) Percentage of the variance of the neuronal firing activity reproduced by each of the first 10 principal components. The inset displays the corresponding cumulative percentage as a function of the considered component. Filled black and shaded red (bar or symbols) correspond to $\tau_{\alpha}$ = 20 ms and $\tau_{\alpha}$ = 2 ms, respectively. Projection of the neuronal response along the first three principal components for b) $\tau_{\alpha}$ = 20 ms and c) $\tau_{\alpha}$ = 2 ms. \red{Each point in the graph correspond to a different time of observation.} The three colors denote the response to the three different inputs, which are constant stimulation currents randomly taken as $I^{(j)} \in [V_{th}, \, V_{th} + \Delta V]$ for $j = 1,2,3$, the experiment is then performed as explained in the text.}

\caption{{\bf Response of the network to an increase in the excitability.} a,f) Network Bursting Rate, and the threshold defined for considering a synchronized event. b,g) Neurons involved in the synchronized events, vertical lines denoted the switching times between the excited $I^{(e)}$ and control $I^{(c)}$ inputs. Colors in the raster indicates the group assigned to the synchronous event using an optimal community partition algorithm. c,h) Synchronized Event Transition Matrix, calculated with a window $T_W = 50$ ms and number of events $S_s = 20$. d,i) Projection of the synchronized events in the 2D space \red{spanned by the first two principal components associated to the covariance matrix of the vectors $W_s$}. e,j) Number of coactive cells in each state. The diagonal elements of the bar graph represents the total number of neurons contributing to one state. Panels (a-e) correspond to $g=8$, while panels (f-j) depict the case $g=1$. In both cases the system is recorded during the time span required to identify $S_s = 20$. Remaining parameters as in Fig. \ref{fig:PonziBenchmark}.}

285 — 1511.06973

\caption{Our proposed framework: given an image, a CNN is first applied to produce the attribute-based representation \textcolor{red}{$\Att(I)$}. The internal textual representation is made up of image captions generated based on the image-attributes. The hidden state of the caption-LSTM after it has generated the last word in each caption is used as its vector representation. These vectors are then aggregated as \textcolor{green}{$\Cap(I)$} with average-pooling. The external knowledge is mined from the KB (in this case DBpedia) and the responses encoded by Doc2Vec, which produces a vector \textcolor{blue}{$\Know(I)$}. The 3 vectors $\mathbf{V}$ are combined into a single representation of scene content, which is input to the VQA LSTM model which interprets the question and generates an answer.}

\caption{Results on the open-answer task for various question types on VQA (July released Version). All results are % in terms of the evaluation metric from the VQA evaluation tools. All results (in \textcolor{gray}{gray}) from \cite{antol2015vqa} are reported based on a different split, which contains 30000 training images and 20000 testing images (3 ground-truth answers per question).}

286 — 1511.07677

\caption{\label{fig:overview} Radial mean flow profiles and corresponding axial wavenumber spectra; %The different methods are visualised in principle: %For each method the upper row of figures shows the underlying flow profile in the duct \mbox{($r$ over $x$)}, while % the bottom row shows the resulting wave number spectrum ($\Im{k_x}$ over $\Re{k_x}$), ({\color{blue}\bf - $\cdot$ -}) acoustic and ({\color{red}\bf - -}) convective wavenumbers. Note that in the XTPP method the convective wavenumber is added to the model. }

287 — 1511.07845

\caption{Predicted symmetry orientations and surface normals for objects segmented out from real world images. The symmetries are shown using the convention in \figref{symSynthRes} and surface normals are mapped into RGB space via the mapping $X\rightarrow$\textcolor{blue}{B}, $Y\rightarrow$\textcolor{red}{R}, $Z\rightarrow$\textcolor{green}{G}.}

\caption{\figlabel{result_normals} Surface normal predictions using our learned and induced models on held out data. Each unit in the figure shows (from left to right) the image, its ground truth surface normals, surface normals predicted by our `Learned' model (trained on all classes) and surface normals predicted by our `Induced' model (trained on the subset of categories not containing this particular class). Surface normals are mapped into RGB space via the mapping $X\rightarrow$\textcolor{blue}{B}, $Y\rightarrow$\textcolor{red}{R}, $Z\rightarrow$\textcolor{green}{G}}

288 — 1511.08589

\caption{On the left is the grid world task with mines (marked as {\color{red}\ding{53}}). On visiting a mine state the agent receives a random reward between $-1$ to $-5$.}

289 — 1511.09349

\caption{Condition numbers in logarithmic scale for observability matrices~$\mathcal{O}_s'$ (\textcolor{blue}{\bf---}) %, $\mathcal{O}_2$~(\textcolor{green!50!black}{\bf---}) and~$\mathcal{O}_s$ (\textcolor{red}{\bf---}) at~$\equ{\ws} = 0$ and~$50\%$ (dotted), $100\%$ (solid) and~$150\%$ (dashed) of nominal flux.}

\caption{Experimental characterization of magnetic saturation on a real $0.75kW$ IM under $5\%$ (\textcolor{blue}{\bf---}), $75\%$ (\textcolor{green}{\bf---}), $100\%$ (\textcolor{red}{\bf---}), $125\%$ (\textcolor{cyan}{\bf---}) and $150\%$ (\textcolor{magenta}{\bf---}) of rated flux.}

\caption{Characterization of magnetic saturation on a simulated $0.75kW$ IM under $5\%$ (\textcolor{blue}{\bf---}), $75\%$ (\textcolor{green}{\bf---}), $100\%$ (\textcolor{red}{\bf---}), $125\%$ (\textcolor{cyan}{\bf---}) and $150\%$ (\textcolor{magenta}{\bf---}) of rated flux.}

290 — 1511.09356

\caption{ \label{fig_momentum_balance_emile} Averaged terms in Eq.~(\ref{eqn_timefiltered_momentum}) for the STBLI case at stations 3 (upper row) and 4 (lower row), filtered using time-scales $ \tau U_\infty / \delta \approx 0.0 $ (left column) and $ \tau U_\infty / \delta \approx 8.5 $ (right column). Pressure-gradient (\solid) , viscous term (\chndot), resolved convection in all three directions (\dashed), and unresolved wall-normal Reynolds stress (\dotted). }

\caption{ \label{fig_momentum_balance_stefan} Averaged terms in Eq.~(\ref{eqn_timefiltered_momentum}) for the APGTBL case at stations 1 to 4 using the time-scale $\tau \rightarrow \infty$. Pressure-gradient (\solid) , viscous term (\chndot), resolved convection in all three directions (\dashed), and unresolved wall-normal Reynolds stress (\dotted). All quantities are given in outer scaling, i.e., scaled by the local boundary-layer edge velocity. }

\caption{ \label{fig_uux_param} Parametrization of the convective term in terms of the pressure-gradient and velocity profile for the APGTBL case at stations 1, 3, 5 and 8 using the time-scale $\tau \rightarrow \infty$. Approximate $\wt{u}_j \partial_j \wt{u}$ from Eq.~(\ref{eqn_uux_parametrized_dpdx}) (\solid), Eq.~(\ref{eqn_uux_parametrized_ux}) (\dashed) , exact $\wt{u} \partial_1 \wt{u}$ (\circle), and exact $\wt{u}_j \partial_j \wt{u}$ (\square).}

\caption[]{ \label{fig_cf} A-priori study: Friction coefficient obtained using various non-equilibrium formulations and compared to existing models. Exact from reference LES (\circle), equilibrium model (\solid), model of \citet{duprat:11} (\dashed), adding $\partial_1 p$ by itself (\chndot), adding $\partial_1 p$ + Eq.~(\ref{eqn_uux_parametrized_dpdx}) with $(y_{pg}^*=4)$ (\solidsquare), adding $\partial_1 p$ + Eq.~(\ref{eqn_uux_parametrized_ux}) (\square). }

291 — 1512.00303

\caption{(\textit{a})Torque on the lid for a water depth of 10 cm: $\square$, measurements; \solidrule: line of best fit. (\textit{b}) friction velocity $u_*$ calculated from torque measurements. }

\caption{Turbulent liquid layer temperature evolution from Experiment 9: \solidrule, measured temperature; \protect\dashedrule , modelled temperature.}

292 — 1512.01476

\caption{\label{fig:dat} (Color online) Statistics of the vertex fluctuations. (a)~Mean square displacement as a function of time in five conditions: control (black), myosin inhibitor (blue), Rho kinase inhibitor (orange), Rho inhibitor (red), and microtubule depolymerization (green). The corresponding best fitting curves are in solid lines (Eq.~(S5)~\cite{supplemental}). The dot dashed line indicates the short time power law, whereas the blue and red dashed lines report the large time behaviors. (b-f)~Distribution of displacement for the five conditions at three times: $5$ ({\color{yellow} $\bullet$}), $25$ ({\color{red}$+$}), and $60$~min ({\color{brown} $\circ$}). Exponential tails appear at long times as a consequence of directed motion events in the vertex dynamics. Results of the simulated dynamics are in solid lines.}

293 — 1512.01825

\caption{\label{vcurve} Velocities of the shifted peaks of the broad H$\alpha$ line and fitted curves. Circle: \cite{Gezari2007}; Down-triangle: \cite{Lewis2010}; Square: \cite{Osterbrock1976}; Up-triangle: \cite{Barr1980}; Diamond: This work; ``x'': \cite{Gaskell1996}; Left-triangle: \cite{Shapovalova2013}; Right-triangle: \cite{Popovic2012}; Plus: \cite{Popovic2014}; Star: SDSS. Solid curves are the best-fit models from the unweighted least squares method (Table~\ref{P_fitting_lsq}). \bluetext{Note that the velocity scale on the vertical axis has been expanded to show detail (i.e., it does not start at zero).} }

\caption{ \label{vcurve2} Velocities of the shifted peaks of the broad H$\alpha$ line and fitted curves. Circle: \cite{Gezari2007}; Down-triangle: \cite{Lewis2010}; Square: \cite{Osterbrock1976}; Up-triangle: \cite{Barr1980}; Diamond: This work; ``x'': \cite{Gaskell1996}; Left-triangle: \cite{Shapovalova2013}; Right-triangle: \cite{Popovic2012}; Plus: \cite{Popovic2014}; Star: SDSS. Solid curves are the best-fit models from the unweighted least squares method (Table~\ref{P_fitting_lsq}). \bluetext{Note that the velocity scale on the vertical axis has been expanded to show detail (i.e., it does not start at zero).} }

294 — 1512.02229

\caption{95\% CL upper limits on cross sections for the heavy Higgs model (HXX) with $m_H = 125$ GeV (a), 200 GeV (b), 400 GeV (c,d) and 1000 GeV (e,f). The colour {\color{red} red} ({\color{blue} blue}) indicates $m_X=20$ GeV (50 GeV) {\it for all curves}. The thin curves in the upper-right corner of all figures show our new $\ETmis$-derived limits on LL particle cross sections for each detector (solid: CMS, dashed: ATLAS). For comparison, the cross section limits from the CMS displaced vertex searches, under the assumption of 100\% branching ratios, are shown by thick curves: displaced leptons searches ($X\to\ell^+\ell^-$) \cite{CMS:2014hka} are indicated by the solid curves for $\ell=e$ and by dashed curves for $\ell=\mu$; whereas displaced jet searches ($X\to q \bar q$) \cite{CMS:2014wda} are indicated by dotted curves. Our new limits are identical in (c) and (d) as well as in (e) and (f) and have been split for clarity. \label{fig:FFX_limit}}

295 — 1512.02252

\caption{$\chi^2$/d.o.f. for different scenarios for the kaon Collins functions: one-parameter (upper panel), two-parameter (central panel) and three-parameter (lower panel) fits. The symbol {\color{red}\textbullet}~means that the corresponding parameter is actually used in the fit, while the symbol {\color{blue}\textopenbullet}~means that the contribution to the Collins asymmetry corresponding to that parameter is not included in the fit. For $N_{\rm fav}^{\rm heavy}$, we explicitly indicate the two different constraints we use: $N_{\rm fav}^{\rm heavy} > 0$ and $N_{\rm fav}^{\rm heavy}<0$. % The symbol ``-" means that a particular choice $\pm|N_{\rm fav}^{\rm heavy}|$ is made for % fitting, where} $|N_{\rm fav}^{\rm heavy}|$ and $-|N_{\rm fav}^{\rm heavy}|$ % mean, respectively, that $N_{\rm fav}^{\rm heavy}$ is assumed to be positive % or negative. \label{fit3}}

296 — 1512.02333

\caption{Gamma-ray luminosity of SNRs (on the left) and PWNe (on the right) compared with the minimal \gray luminosity required to produce detectable HE \nray flux. The filled area corresponds to different power-law indicis of initial protons ($\alpha=(1.5\div3.5)$).}

297 — 1512.02415

\caption{(a) Stacked front-view images (taken from the laser beam axis) of the expanding drop shown in Fig.~\ref{fig1}c. The dashed lines are the ellipses best fitting the contours, from which the equivalent radii $R$ are determined. (b) Relative expansion of the drop obtained from the views shown in (a) (\protect\includegraphics{figure13_legend01.pdf}) and from later times (\protect \includegraphics{figure13_legend02.pdf}). The linear fit (solid line) to the first points yields a dimensional expansion speed $\dot{R} = 14.6\, R_0/\tau_{\rm c}$ = 4.2\,m/s.}

298 — 1512.02497

\caption{ {\bf System overview.} Our system takes as input individual 2D image object proposal windows (top-left) generated by the selective search algorithm~\cite{Uijlings13}. The image window is passed through the initial layers of a pre-trained CaffeNet model~\cite{jia2014caffe} to generate a feature vector (top-middle). Here, we visualize CNN features using the inversion network of \cite{Dosovitskiy15} \textcolor{red}{(outlined in red)}, which infers the original image given a CNN layer's response. In an offline step (bottom-left), we similarly pass rendered views of a library of 3D object CAD models through the initial layers of CaffeNet and record their responses. As there is a domain gap between the appearance of natural images and rendered views of CAD models, we learn to adapt the features for a natural image to better align to those of CAD models (top-right). We compare the features and return the view that best matches the style and pose of the input image (bottom-right). }

299 — 1512.02505

\caption{(a-c) Different colorings of the same graph: (a)~a traditional $4$-coloring, (b)~an \konecoloring{3} (c)~a \twocoloring; (d)~a star with three leaves; its \emph{center} has degree~$3$.}

\caption{(a)~An outerplanar graph that is not \twocolorable. (b)~An outerpath, whose spine edges are drawn as dashed segments. Dotted arcs highlighted in gray correspond to edges belonging to the fan of each spine vertex. Note that $|f_6|=0$.}

\caption{Illustration of: \sref{fig:k24}~a graph with $6$ vertices, \sref{fig:chain}~a chain of length $3$, \sref{fig:reduction1}~the reduction from \naesat to \twocoloring: $\phi = (x_1 \lor x_2 \lor x_3) \land (\neg x_1 \lor \neg x_2 \lor \neg x_3)$. The solution corresponds to the assignment $x_1=false$ and $x_2=x_3=false$. Sets $O_{x_1}$, $E_{x_2}$ and $E_{x_3}$ ($E_{x_1}$, $O_{x_2}$ and $O_{x_3}$, resp.) are colored gray (white, resp.).}

\caption{ (a)~The complete graph on six vertices $K_6$. (b)~The attachment-graph for the planar case. (c)~A planar graph of max-degree~$4$ that is not \twocolorable.}

300 — 1512.02684

\caption{(a) Porcine experimental set-up (b) Relay position changes (c) Voronoi region samples of relay (*) positions; \textcolor{red}{$\triangle$} - implants; \textcolor{blue}{ $\square$} - surface nodes }

301 — 1512.03062

\caption{\SystemADES. The 1D spectra for A1 and A2 are shown on the left. There are strong emission lines at the same observed wavelength $4535$\AA\in both spectra. Taking into account the absence of other spectral features in the R150 spectra, along with the photometric redshift of the lens galaxy, we assign the features to be\lymanalpha, which gives redshifts of $\zsource=\SystemAspeczA$ and $\SystemAspeczB$ for A1 and A2, respectively. The orientation of the slits (red boxes) is shown in the upper right DES $g$-band image. In the color coadd image in the middle right panel, the two features of interest are labeled ``A1'' and ``A2'', the Einstein radius is marked by a white arc, and the small scale bar shows the size of the image. In the lower right is an identical color coadd image. All images are oriented north up, east left. The flat, zero-count portions of the spectra near $\lambda \sim 4200$\AA\and$\sim5300$\AA\show the chip gaps in the detectors: we didn't take an offset observation to cover the chip gaps for this system.}

\caption{\SystemEDES. The 1D spectra for B1 and B2 are shown on the left. The spectrum for \arcA\is not shown, but has an emission line consistent with an interpretation as a foreground object. Prominent emission lines in both arcs at the same observed wavelength of 5117\AA. We conclude that this is \lymanalpha\emission, from which we obtain redshifts of$\zsource=\SystemEspeczB$ for \arcB\and$\SystemEspeczC$ for \arcC. The object labeled \arcA\also has emission lines (not shown), which we interpret as\OII\and\OIII. This gives a redshift of $\zsource = \SystemEspeczA$, making object \arcA\a foreground object. The orientation of the slits (red boxes) is shown in the upper right DES$g$-band image. In the color coadd image in the middle right panel the three features of interest are labeled ``A'', ``B1'' and ``B2'', the Einstein radius is marked by a white arc, and the small scale bar shows the size of the image. In the lower right is an identical color coadd image. All images are oriented north up, east left.}

\caption{\redmapper\Group and Cluster Properties\label{table:clusterproperties}}

302 — 1512.03066

\caption{Damping of Alfv\'enic turbulence in low beta partially ionized gas. The damping of turbulence happens when the rate of damping (solid line) intersects the dashed line corresponding to the cascading rate. From\cite{Xu_etal:2015}.}

\caption{Candidate Events for Turbulent Reconnection. {\it MHD Turbulence Simulation (Top Panels) and High-Speed Solar Wind (Bottom Panels)}. The left panels show magnetic field components and the right panels velocity components, both rotated into a local minimum-variance frame of the magnetic field. The component of maximum variance in \textcolor{red}{{\bf red}} is the apparent reconnecting component, the component of medium variance in \textcolor{green}{green} is the nominal guide-field direction, and the minimum-variance direction in \textcolor{blue}{blue} is perpendicular to the reconnection layer. Reprinted figure with permission from \cite{Lalescu_etal:2015}. Copyright (2015) by the American Physical Society. \label{Gosling-criterion}}

303 — 1512.03584

\caption{\small Comparison of the numerical simulation ({\color{blue} ---}) with the analytical second order rational soliton ({\color{red} -.-}) for the rogue wave profile.}

\caption{\small Histogram of the maxima appearing in simulations of the chaotic wave field for $\alpha=0.2$, $\mu=0$ ({\color{blue} ---}) vs. $\mu=0.8$ ({\color{red} -.-})}

\caption{\small Histogram of the maxima appearing in simulations of the chaotic wave field for $\mu=0$, $\alpha=0.2$ ({\color{blue} ---}) vs. $\alpha=0.6$ ({\color{red} -.-})}

304 — 1512.04285

\caption{Single-frame excerpts from video recordings of metallic objects concealed by opaque plastic tape. (a) Utility blade (\textcolor{blue}{Media~1}). (b) Dentist's pick (\textcolor{blue}{Media~2}). (c) Paper clip (\textcolor{blue}{Media~3}). (d) Plastic/wire tie twisted into the shape of a loop (\textcolor{blue}{Media~4}). (Sample figure adapted from \cite{samplefig}.)}

\caption{Single-frame excerpts from video recordings of metallic objects concealed by opaque plastic tape. (a) Utility blade (\textcolor{blue}{Media~1}). (b) Dentist's pick (\textcolor{blue}{Media~2}). (c) Paper clip (\textcolor{blue}{Media~3}). (d) Plastic/wire tie twisted into the shape of a loop (\textcolor{blue}{Media~4}). (Sample figure adapted from \cite{samplefig}.)}

305 — 1512.04468

\caption[]{Density function (left column) and point-wise error (right column) compared to $10^6$ SSA trajectories. In the left column, the method presented above is shown in light red, while SSA is shown in purple. The top figures denote larger values of the error control parameter $\epsilon$, while the lower figures show smaller values. $\rho$ denotes the ratio of gamma distributed random variables to exponential random variables. Note that the Monte Carlo sampling error is $\mathcal{O}\left( 10^{-3} \right)$. } \label{fig:PDFerror} \end{center} \end{figure} \begin{figure}[htp] \begin{center} \includegraphics[width= 0.75\textwidth] {ConvergencePlot.jpg} \caption[]{Error of the exit time method: $l^1$ norm in red, and $l^2$ norm in blue, $\mathcal{O}(\epsilon)$ dashed and $\mathcal{O}\left(\epsilon^2 \right)$ solid lines. Note that the Monte Carlo error is $\mathcal{O}\left( 10^{-3} \right)$. } \label{fig:convergence} \end{center} \end{figure} \begin{figure}[htp] \begin{center} \includegraphics[width= 0.75\textwidth] {EfficiencyTest.jpg} \caption[]{Random variable generation speedup: Speedup of gamma distributed samples compared to exponentially distributed. In the SIR example, $\rho$ ranges from $0.01 - 0.03$, yielding a speedup of $S(\rho)$ is between $10-30$. Note that this is the speedup of the random number generation not the total simulation time.} \label{fig:speedup} \end{center} \end{figure} Since increasing $\epsilon$ will increase the error, we compared the density function of the exit times from the method presented here and the classical simulation algorithm in Figure \ref{fig:PDFerror}. We have shown the two distributions in the left panel as well as the difference between them on the right. We note that the exit time method is able to capture the correct distribution. In addition, we reported the value for $\rho \triangleq N(\Gamma)/N(E)$ which is the number of gamma distributed random variables needed for the exit time method divided by the number of exponential random variables needed for the stochastic simulation algorithm. In Figure \ref{fig:convergence} we have plotted convergence with respect to $\epsilon$, which is in accord with the error analysis included in the derivation in section 2. We note that there is a tapering off in the error which is incident to the Monte Carlo error becoming larger. In Figure \ref{fig:speedup} we have shown the speedup where $\rho$ is the ratio of the number of gamma random variates to the number of exponential random variates. We note that this is not the speedup of the simulation, but merely for the computation of random numbers. \section{Conclusion} A original method was presented to compute the exit time for the stochastic simulation algorithm. The method was based on the addition of a series of random variables and was derived using the convolution theorem. We derived the final distribution and showed one approximation method. The result led to a formulation of the stochastic simulation algorithm that requires fewer random variates. As shown in the results section above for a typical nonlinear model, the error control parameter $\epsilon$ can be suitably chosen such that the number of random variates needed to resolve the exit time is reduced. Equation \eqref{eq:finalSample} has better convergence properties than leaping algorithms since the error is second order whereas leaping algorithms are typically first order, therefore, the Monte Carlo error will still dominate the total error. While the method is similar to R-Leaping \cite{Auger:2006} in that the distribution for time is drawn from a Gamma distribution, it differs in that the propensities are grouped according to their magnitude and the trajectory for the state transitions is exact. Although the number of random variates has been reduced, the bottleneck of the stochastic simulation algorithm is still the re-computation of the propensities at each time-step. While we report relatively modest speed-up compared to the classical stochastic simulation algorithm, the derivation and application of the method may be beneficial to other areas of algorithmic research. Indeed the derivation is not limited to the stochastic simulation algorithm and in principle could be used in other Monte Carlo algorithms. If $Y_{i}$ is a sequence of random variables for $i = 1, \ldots, n$ such that $Y_{i} \sim \mathcal{D}(\Theta_i)$, where $\Theta_i$ is independent of $Y_1, \ldots, Y_n$, then the derivation could be used to find an expression for $Y_1 + \ldots + Y_n$. \section{Acknowledgements} The author thanks Bill and Melinda Gates for their active support of this work and their sponsorship through the Global Good Fund. \appendix \section{Appendix} \label{sec:Appendix} This appendix is provided for completeness and includes derivations of mathematical results used in the exit time method presented in section 2. \subsection{Hypoexponential Distribution} \label{subsec:HypoexponentialDistribution} \subsubsection*{Derivation of the Probability Density Function} \label{subusb:DerivationOfHypoexponentialDistribution} We begin by attempting to rewrite the product in terms of partial fractions: \begin{eqnarray} \prod_{i=1}^n \frac{\lambda_i}{\lambda_i + s} \stackrel{\text{!}}{=} \sum_{i=1}^n \frac{C_i}{\lambda_i + s } \end{eqnarray} where $C_i \triangleq C_i(\lambda_1, \ldots, \lambda_n)$ must be determined and $ \stackrel{\text{!}}{=}$ denotes `shall be equal to'. Then, \begin{eqnarray} \prod_{i=1}^n \frac{\lambda_i}{\lambda_i + s} &=& \sum_{i=1}^n \left( \frac{C_i }{\left(\lambda_i + s\right)} \frac{ \prod_{\substack{j=1\\j \neq i}}^n \left( \lambda_j + s \right)}{ \prod_{\substack{j=1\\j \neq i}}^n \left( \lambda_j + s \right)} \right) \implies \\ \prod_{i=1}^n \lambda_i &=& \sum_{i=1}^n \left(C_i \prod_{\substack{j=1\\j \neq i}}^n \left( \lambda_j + s \right) \right), \end{eqnarray} which holds $\forall s$. If $\mathfrak{Re}(s) = -\lambda_1$ and $\mathfrak{Im}(s) = 0$, then \begin{eqnarray} \prod_{i=1}^n \lambda_i &=& C_1 \prod_{\substack{j=1\\j \neq 1}}^n \left( \lambda_j - \lambda_1 \right) + \underbrace{C_2 \prod_{\substack{j=1\\j \neq 2}}^n \left( \lambda_j - \lambda_1 \right)}_{= 0} + \underbrace{\ldots}_{=0} \\ C_1 &=& \frac{ \prod_{i=1}^n \lambda_i}{ \prod_{\substack{j=1\\j \neq 1}}^n \left( \lambda_j - \lambda_1 \right) } \\ &=& \lambda_1 \prod_{\substack{j=1\\j \neq 1}}^n \frac{\lambda_j}{\lambda_j - \lambda_1} \end{eqnarray} therefore \begin{eqnarray} C_i = \lambda_i \prod_{\substack{j=1\\j \neq i}}^n \frac{\lambda_j}{\lambda_j - \lambda_i} \end{eqnarray} \begin{eqnarray} l_i \triangleq l_i\left(\lambda_1, \ldots, \lambda_n \right) \triangleq \prod_{\substack{j=1\\j \neq i}}^n \frac{\lambda_j}{\lambda_j - \lambda_i} \end{eqnarray} then \begin{eqnarray} \prod_{i=1}^n \frac{\lambda_i}{\lambda_i + s} &=& \sum_{i=1}^n l_i \frac{\lambda_i}{\lambda_i + s}. \end{eqnarray} In order to find the analytical form of the original convolution, we apply the inverse Laplace transform, viz.: \begin{eqnarray} \mathcal{L}^{-1}\left\{ \prod_{i=1}^n \frac{\lambda_i}{\lambda_i + s} \right\}(t) &=& \mathcal{L}^{-1}\left\{ \sum_{i=1}^n l_i \frac{\lambda_i}{\lambda_i + s} \right\}(t) \\ &=& \sum_{i=1}^n l_i \mathcal{L}^{-1}\left\{\frac{\lambda_i}{\lambda_i + s} \right\}(t) \end{eqnarray} The inverse can be found by noting that $ \mathcal{L}^{-1}\left\{\mathcal{L}\left\{f(t) \right\}(s) \right\}(t) = f(t)$ and therefore $\mathcal{L}^{-1}\left\{\frac{\lambda_i}{\lambda_i + s} \right\}(t) = \lambda_i e^{-\lambda_i t}$. \subsubsection*{Summary (see \cite{Bolch:2006}):} \begin{eqnarray} \mathcal{L}^{-1}\left\{ \prod_{i=1}^n \frac{\lambda_i}{\lambda_i + s} \right\}(t) = \sum_{i=1}^n \left( \prod_{\substack{j=1\\j \neq i}}^n \frac{\lambda_j}{\lambda_j - \lambda_i} \right) \lambda_i e^{-\lambda_i t}, \end{eqnarray} \subsubsection*{Drawing Random Variates from the Probability Density Function} We use the inversion theorem \cite{Devroye:1986} to obtain: \label{sec:RandomVariatesFromPDF} \begin{eqnarray} p(t; \lambda_1, \ldots, \lambda_n) &=& \sum_{i=1}^n l_i \lambda_i e^{-\lambda_i t} \\ \int_0^{t} p(t'; \lambda_1, \ldots, \lambda_n)~\text{d}t' &=& \int_0^{t} \sum_{i=1}^n l_i \lambda_i e^{-\lambda_i t'} ~\text{d}t' \\ &=& \sum_{i=1}^n l_i \int_0^{t} \lambda_i e^{-\lambda_i t'} ~\text{d}t' \\ &=& \sum_{i=1}^n l_i \left( 1-e^{-\lambda_i t} \right) \end{eqnarray} Define the cumulative distribution $\hat{p}(t; \lambda_1, \ldots, \lambda_n) = \sum_{i=1}^n l_i \left(1-e^{-\lambda_i t} \right)$, then, find $t$ such that $ \sum_{i=1}^n l_i \left(1-e^{-\lambda_i t}\right) = r$, where $r \sim \mathcal{U}(0,1)$ by a root finding method. \subsection{Frequency-Domain of Gamma/Erlang Distribution} \label{sec:GammaFrequency} Define the gamma distribution: \begin{eqnarray} q(t; \lambda, n) \triangleq \frac{\lambda^n t^{n-1} e^{-\lambda t}}{\Gamma(n)}, \end{eqnarray} where $t \in [0, \infty)$, $\lambda > 0$, $n >0$, and $\Gamma(n) = (n-1)!$. Laplace's transform yields (`i.b.p.' denotes integration by parts): \begin{eqnarray} \mathcal{L}\left\{q(t; \lambda, n) \right\}(s) &=& \int_0^{\infty} \frac{\lambda^n t^{n-1} e^{-\lambda t}}{\Gamma(n)} e^{-s t}~\text{d}t \\ &=& \frac{\lambda^n}{\Gamma(n)} \int_0^{\infty} t^{n-1} e^{-t (\lambda+s)}~\text{d}t \\ &\stackrel{\text{i.b.p}}{=}& \frac{\lambda^n}{\Gamma(n)} \int_0^{\infty} (n-1) t^{n-2} \frac{1}{(\lambda + s)} e^{-t (\lambda+s)}~\text{d}t \\ &=& \frac{\lambda^n}{(\lambda + s)}\frac{1}{\Gamma(n)} (n-1) \int_0^{\infty} t^{n-2} e^{-t (\lambda+s)}~\text{d}t \\ &\stackrel{\text{i.b.p}}{=}& \frac{\lambda^n}{(\lambda + s)^2}\frac{1}{\Gamma(n)} (n-1) (n-2) \int_0^{\infty} t^{n-3} e^{-t (\lambda+s)}~\text{d}t~~~~ \\ &\stackrel{\text{i.b.p}}{=}& \ldots \\ &=& \frac{\lambda^n}{(\lambda + s)^n}\frac{1}{\Gamma(n)} (n-1)! \\ %&=& \frac{\lambda^n}{(\lambda + s)^n}\frac{1}{\Gamma(n)} \Gamma(n) \\ &=& \left( \frac{\lambda}{\lambda + s} \right)^n \end{eqnarray} \subsubsection*{Summary:} \begin{eqnarray} \mathcal{L}\left\{\frac{\lambda^n t^{n-1} e^{-\lambda t}}{\Gamma(n)} \right\}(s) = \left( \frac{\lambda}{\lambda + s} \right)^n \end{eqnarray} \subsection{The Residue Theorem} The residue theorem \cite{residueMathworld} can be used to compute the inverse Laplace transform: \label{sec:residueTheory} \begin{eqnarray} f(t) = \mathcal{L}^{-1}\left\{\tilde{f}(s) \right\}(t) = \sum_k \text{Res} \left(\tilde{f}(s) e^{s t}, s_k \right), \end{eqnarray} where $s_k$ is a pole and the complex residue for a pole $s_k$ of order $\eta$ is \begin{eqnarray} \text{Res} \left(\tilde{f}(s) e^{s t}, s_k \right)= \frac{1}{(\eta-1)!} \lim_{s \rightarrow s_k} \frac{\text{d}^{\eta-1}}{\text{d}s^{\eta-1}} \left((s-s_k)^{\eta} \tilde{f}(s) e^{s t} \right)\end{eqnarray} \subsection{Approximating the Density Function} \label{sec:Approx} Let \begin{eqnarray} f(\xi, s) \triangleq \frac{\xi}{\xi + s} \end{eqnarray} and let $\lambda_1 \triangleq \lambda + \epsilon$ and $\lambda_2 \triangleq \lambda - \epsilon$, where $\epsilon \ll 1$. Then, we can write \begin{eqnarray} f(\lambda_1, s) f(\lambda_2, s) = f(\lambda, s)^2 + \mathcal{O}\left(\epsilon^2 \right). \end{eqnarray} This can be seen by expanding $f(\lambda_1, s)$ and $f(\lambda_2, s)$: \begin{eqnarray} f(\lambda_1, s) = f(\lambda + \epsilon, s) = f(\lambda, s) + \epsilon \frac{\partial f}{\partial \lambda} + \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2} + \mathcal{O}\left(\epsilon^3\right). \end{eqnarray} and \begin{eqnarray} f(\lambda_2, s) = f(\lambda - \epsilon, s) = f(\lambda, s) - \epsilon \frac{\partial f}{\partial \lambda} + \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2} - \mathcal{O}\left(\epsilon^3 \right). \end{eqnarray} then \begin{eqnarray} f(\lambda_1, s) f(\lambda_2, s) = \left(f(\lambda, s) + \epsilon \frac{\partial f}{\partial \lambda} + \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2} + \mathcal{O}\left(\epsilon^3 \right) \right)\times \\ \left(f(\lambda, s) - \epsilon \frac{\partial f}{\partial \lambda} + \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2} - \mathcal{O}\left(\epsilon^3\right)\right)~~~\\ = f(\lambda, s)^2 + \epsilon \frac{\partial f}{\partial \lambda} f(\lambda, s) + \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2} f(\lambda, s) - \epsilon \frac{\partial f}{\partial \lambda} f(\lambda, s) \\ - \epsilon^2 \left( \frac{\partial f}{\partial \lambda} \right)^2 \frac{\epsilon^2}{2} \frac{\partial^2 f}{\partial \lambda^2}f(\lambda, s) + \mathcal{O}\left(\epsilon^3\right)\\ = f(\lambda, s)^2 + \mathcal{O}\left(\epsilon^2\right)\end{eqnarray} \subsubsection*{Summary:} \begin{eqnarray} \left(\frac{\lambda_1}{\lambda_1 + s}\right)\left(\frac{\lambda_2}{\lambda_2 + s}\right)= \left(\frac{\lambda}{\lambda + s} \right)^2 + \mathcal{O}\left(\epsilon^2 \right), \end{eqnarray} where $\lambda_1 \triangleq \lambda + \epsilon$ and $\lambda_2 \triangleq \lambda - \epsilon$. \subsection{Determining the $\tilde{\lambda}_k$s and $n_k$s for a given $\epsilon$} \label{sec:Epsilon} We will partition the set $\Lambda = \{\lambda_1, \ldots, \lambda_n \}$ into $m$ disjoint sets $A_1, \ldots, A_m$, i.e. $\bigcup_{k=1}^m A_k = \Lambda$ and $\bigcap_{k=1}^K A_k = \emptyset$. Initially set $A_k = \emptyset$ $\forall k$ and let $\hat{\Lambda}$ be the set $\Lambda$ sorted in descending order. The procedure is as follows: for $k = 1, \ldots, m$, set $A_k = \{ a | a \in \hat{\Lambda}~\wedge~a \ge \hat{\Lambda}_1 - \epsilon \hat{\Lambda}_1 \}$, where $\hat{\Lambda}_1$ is the first element in the set $\hat{\Lambda}$, and then set $\hat{\Lambda} =\hat{\Lambda} - \bigcup_{j=1}^{k} A_j$ and repeat until $\hat{\Lambda} = \emptyset$. Note that $m$ need not be known a priori and that $\epsilon = 0$ is valid. For each set $A_k$, calculate the mean rate $\tilde{\lambda}_k$: \begin{eqnarray} \frac{1}{\tilde{\lambda}_k} = \frac{1}{|A_k|}\sum_{i \in A_k} \frac{1}{\lambda_i} \end{eqnarray} and set $n_k = |A_k|$, which is to be used in equation \ref{eq:finalSample}, where $|.|$ denotes the number of elements of the set. %We will partition the set $\Lambda = \{\lambda_1, \ldots, \lambda_n \}$ into $m$ disjoint sets $A_1, \ldots, A_m$, i.e. $\bigcup_{k=1}^m A_k = \Lambda$ and $\bigcap_{k=1}^K A_k = \emptyset$. % %Initially set $A_k = \emptyset$ $\forall k$ and let $\hat{\Lambda}$ be the set $\Lambda$ sorted in descending order. The procedure is as follows: for $k = 1, \ldots, m$, set $A_k = \{ a | a \in \hat{\Lambda}~\wedge~a \ge \hat{\Lambda}_1 - \epsilon \hat{\Lambda}_1 \}$, where $\hat{\Lambda}_1$ is the first element in the set $\hat{\Lambda}$, and then set $\hat{\Lambda} =\hat{\Lambda} - \bigcup_{k=1}^{j} S_k$ and repeat until $\hat{\Lambda} = \emptyset$. Note that $m$ need not be known a priori and that $\epsilon = 0$ is valid. % %For each set $A_k$, calculate the mean rate $\tilde{\lambda}_k$: %\begin{eqnarray} %\frac{1}{\tilde{\lambda}_k} = \frac{1}{|A_k|}\sum_{i \in A_k} \frac{1}{\lambda_i} %\end{eqnarray} %and set $n_k = |A_k|$, where $|.|$ denotes the number of elements of the set. %\begin{table} % \begin{tabular}{|c| c | c | c | c| } %\hline % $\epsilon$ & $ 0.5$ & $ 0.25$ & $0.125$ & $0$ \\ \hline % Running Time [s] & 4.870 & 4.889 & 4.939 & 5.977\\ % \hline % \end{tabular} %\caption[Table caption text]{Running time shown in seconds, $\epsilon=0$ denotes the standard simulation algorithm.} %\end{table} \begin{thebibliography}{17} \expandafter\ifx\csname natexlab\endcsname\relax\def{\}{natexlab}#1{#1}\fi \expandafter\ifx\csname bibnamefont\endcsname\relax \def{\}{bibnamefont}#1{#1}\fi \expandafter\ifx\csname bibfnamefont\endcsname\relax \def{\}{bibfnamefont}#1{#1}\fi \expandafter\ifx\csname citenamefont\endcsname\relax \def{\}{citenamefont}#1{#1}\fi \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[{\citenamefont{Chandrasekhar}(1943)}]{Chandrasekhar:1943} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Chandrasekhar}}, \bibinfo{journal}{Rev. Mod. Phys.} \textbf{\bibinfo{volume}{15}}, \bibinfo{pages}{0001} (\bibinfo{year}{1943}), ISSN \bibinfo{issn}{0034-6861}. \bibitem[{\citenamefont{Dietz}(1967)}]{Dietz:1967} \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Dietz}}, \bibinfo{journal}{J. Roy. Stat. Soc. A} \textbf{\bibinfo{volume}{130}}, \bibinfo{pages}{505} (\bibinfo{year}{1967}). \bibitem[{\citenamefont{Kampen}(2007)}]{Kampen:2007} \bibinfo{author}{\bibfnamefont{N.~V.} \bibnamefont{Kampen}}, \emph{\bibinfo{title}{Stochastic Processes in Physics and Chemistry}} (\bibinfo{publisher}{North Holland}, \bibinfo{year}{2007}), \bibinfo{edition}{3rd} ed. \bibitem[{\citenamefont{Gardiner}(2009)}]{Gardiner:2009} \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gardiner}}, \emph{\bibinfo{title}{Stochastic Methods: A Handbook for the Natural and Social Sciences}} (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{year}{2009}), \bibinfo{edition}{4th} ed. \bibitem[{\citenamefont{Gillespie}(1976)}]{Gillespie:1976} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gillespie}}, \bibinfo{journal}{J. Comput. Phys.} \textbf{\bibinfo{volume}{22}}, \bibinfo{pages}{403} (\bibinfo{year}{1976}). \bibitem[{\citenamefont{Gillespie}(1977)}]{Gillespie:1977} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gillespie}}, \bibinfo{journal}{J. Phys. Chem.} \textbf{\bibinfo{volume}{81}}, \bibinfo{pages}{2340} (\bibinfo{year}{1977}). \bibitem[{\citenamefont{Gillespie}(2009)}]{Gillespie:2009} \bibinfo{author}{\bibfnamefont{D.~T.} \bibnamefont{Gillespie}}, \bibinfo{journal}{J. Phys. Chem. B} \textbf{\bibinfo{volume}{113}}, \bibinfo{pages}{1640} (\bibinfo{year}{2009}), ISSN \bibinfo{issn}{1520-6106}. \bibitem[{\citenamefont{Gillespie}(2001)}]{Gillespie:2001} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gillespie}}, \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{115}}, \bibinfo{pages}{1716} (\bibinfo{year}{2001}). \bibitem[{\citenamefont{Auger et~al.}(2006)\citenamefont{Auger, Chatelain, and Koumoutsakos}}]{Auger:2006} \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Auger}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Chatelain}}, \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Koumoutsakos}}, \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{125}}, \bibinfo{pages}{084103} (\bibinfo{year}{2006}), ISSN \bibinfo{issn}{0021-9606}. \bibitem[{\citenamefont{Cao et~al.}(2006)\citenamefont{Cao, Gillespie, and Petzold}}]{Cao:2006} \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Cao}}, \bibinfo{author}{\bibfnamefont{D.~T.} \bibnamefont{Gillespie}}, \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{Petzold}}, \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{124}}, \bibinfo{pages}{044109} (\bibinfo{year}{2006}), ISSN \bibinfo{issn}{0021-9606}. \bibitem[{\citenamefont{Anderson et~al.}(2011)\citenamefont{Anderson, Ganguly, and Kurtz}}]{Anderson:2011} \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Anderson}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ganguly}}, \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kurtz}}, \bibinfo{journal}{Journal of Applied Probability} \textbf{\bibinfo{volume}{21}}, \bibinfo{pages}{2226} (\bibinfo{year}{2011}), ISSN \bibinfo{issn}{0021-9606}. \bibitem[{\citenamefont{Devroye}(1986)}]{Devroye:1986} \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Devroye}}, \emph{\bibinfo{title}{Non-Uniform Random Variate Generation}} (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{year}{1986}). \bibitem[{\citenamefont{Springer}(1979)}]{Springer:1979} \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Springer}}, \emph{\bibinfo{title}{The Algebra of Random Variables}} (\bibinfo{publisher}{Wiley}, \bibinfo{year}{1979}). \bibitem[{\citenamefont{Canosa}(1969)}]{Canosa:1969} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Canosa}}, \bibinfo{journal}{J. Math. Phys.} (\bibinfo{year}{1969}). \bibitem[{\citenamefont{Kermack and McKendrick}(1927)}]{Kermack:1927} \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Kermack}} \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{McKendrick}}, \bibinfo{journal}{Proc. R. Soc. Lond. A.} \textbf{\bibinfo{volume}{115}}, \bibinfo{pages}{700} (\bibinfo{year}{1927}), ISSN \bibinfo{issn}{1463-9076}. \bibitem[{\citenamefont{Bolch}(2006)}]{Bolch:2006} \bibinfo{author}{\bibfnamefont{S.~d. M. H. T. K.~S.} \bibnamefont{Bolch}, \bibfnamefont{Gunter;~Greiner}}, \emph{\bibinfo{title}{Queueing Networks and Markov Chains: Modeling and Performance Evaluation with Computer Science Applications}} (\bibinfo{publisher}{Wiley-Blackwell}, \bibinfo{year}{2006}). \bibitem[{\citenamefont{Rowland and Weisstein}()}]{residueMathworld} \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Rowland}} \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.~W.} \bibnamefont{Weisstein}}, \emph{\bibinfo{title}{Complex residue. from mathworld--a wolfram web resource.}}, \urlprefix\url{http://mathworld.wolfram.com/ComplexResidue.html}. \end{thebibliography} \end{document}}

306 — 1512.04537

\caption{The dependence of the mean surface density of star formation, $\csigmasfr$, on star formation rate and redshift. {\bf Panel a:} $\csigmasfr$ of \eagle\galaxies with stellar mass$>10^8$ M$_\odot$, as a function of their 1500\AA\UV luminosity ($L_{1500}$, bottom axis, corresponding star formation rate, $\dot M_\star$, top axis), with coloured lines depicting the median relation, and coloured regions enclosing the 25th to 75th percentile range. Different colours refer to different redshifts ({\color{blue}{\em blue}}, {\color{blue} $z=0$}, {\color{green}{\em green}}, {\color{green} $z=3$}, {\color{red}{\em red}}, {\color{red} $z=6$}, {\color{magenta}{\em magenta}}, {\color{magenta} $z=8$}). $\csigmasfr$ increases with $\dot M_\star$ at given $z$, and with $z$ at given $M_\star$. Horizontal black line shows the critical star formation rate surface density ($\sigmasfrcrit$) for outflows \citep{Heckman01}. {\bf Panel b:} Number density of \eagle\galaxies in bins of$\csigmasfr$, using the same colour scheme. At $z=0$, galaxies with $\csigmasfr\gtrsim\sigmasfrcrit$ above which significant winds develop are present in the simulation, but they are very rare. However at $z\gtrsim 6$, most \eagle\galaxies with$\dot M_\star>1$~M$_\odot$~yr$^{-1}$ are above this limit, and are expected to drive strong winds. Observed values of $\csigmasfr$ are plotted in panel a, with colours depending on $z$ matching those of the simulation; the region of low $\csigmasfr$ where winds are not expected is hashed. The data shown are: $z=0$, {\em Sloan Digital Sky Survey} measurements \protect\citep{Brisbin12} ({\color{blue} squares}), the galaxies with high values of $f_{\rm esc}$ from \protect\cite{Borthakur14} ({\color{blue} cross}) and \protect\cite{Izotov16} ({\color{blue} plus}), Lyman-break galaxies from \citet{Giavalisco96} at $z=3$ ({\color{green} stars}), from \citet{Curtis14} at $z=6$ ({\color{red} small dots}) and at $z=8$ ({\color{magenta} triangles}). \eagle\galaxies reproduce the dependence of$\csigmasfr$ on $\dot M_\star$ and $z$.}

307 — 1512.05411

\caption{\mis\denotes maximal independent set,\mm\denotes maximal matching,\dcolor\denotes$\Delta+1$ vertex colouring, $(1-\eps)$-\mcm\denotes$(1-\eps)$-approximated maximum cardinality matching, and $(1-\eps)$-\mwm\denotes$(1-\eps)$-approximated maximum weighted matching. All the upper bounds presented in this table are of algorithms which are deterministic and stateless. All the upper bounds are presented under the assumption that $\Delta = O(1)$ and $\eps = O(1)$. For weighted graphs, the ratio between the maximum to minimum edge weight is denoted by $\Gamma$. }

308 — 1512.05869

\caption{Phase of the roundtrip matrix element given in Eq. (\ref{eqn:roundtrip}) for different cavity-PTE distances. Each point has been color coded according to the parity of the numerically exact transmission spectrum (obtained from Eqs. (\ref{eqn:Trans_mat_eqn}-\ref{eqn:RT_mat_eqn})), such that {\color{red}red} parity structures are marked by {\color{red}red} crosses and {\color{blue}blue} parity structures are marked with {\color{blue}blue} dots. The black circle indicates the chosen cavity-PTE distance used for Fig. \ref{fig:Trans_different_RTMin}.}

\caption{Transmission spectrum for $\Phi_{RT}(\omega_\mathrm{min}) = \pi/4$ ({\color{blue} blue} solid), $\Phi_{RT}(\omega_\mathrm{min}) = 0$ ({\color{black} black} dot-dashed) and $\Phi_{RT}(\omega_\mathrm{min}) = -\pi/4$ ({\color{red} red} dashed), where all other parameters are those for $d = 6.08a$.}

309 — 1512.06689

\caption{\color[HTML]{0000FF}{\label{fig:1} An illustration of the two labs gedanken experiment with free agents.}}

\caption{\color[HTML]{0000FF}{\label{fig:2} An illustration of the two machines gedanken experiment. The paradox is symmetric, but for simplicity it is shown to reside on the B side. }}

310 — 1512.07280

\caption{The allowed regions in ($\lambda_{1},\lambda_{4}$) plans after imposing theoretical and experimental constraints. (\textcolor{cyan}{cyan}) : Excluded by $\mu$ constraints, (\textcolor{red}{red}) : Excluded by $\mu$+Unitarity constraints, (\textcolor{green}{green}) : Excluded by $\mu$+Unitarity+BFB constraints, (\textcolor{blue}{blue}) : Excluded by $\mu$+Unitarity+BFB+$R_{\gamma\gamma}$ constraints, (\textcolor{yellow}{yellow}) : Excluded by $\mu$+Unitarity+BFB $R_{\gamma\gamma}$\&$T_d=0$ $\land$ $T_t=0$ constraints. Only the brown area obeys ALL constraints. Our inputs are $\lambda = 0.52$, $-2 \le \lambda_1 \le 12$, $\lambda_2 = -\frac{1}{6}$, $\lambda_3 = \frac{3}{8}$, $-12 \le \lambda_4 \le 2$, $v_t = 1$ GeV and $\mu = 1$ GeV.}

311 — 1512.07286

\caption{\label{figchaos}{\bf Regions of chaos and quasiperiodicity for the Standard Map.} Here $r = 1.4$ and $\QC_N$ is calculated with $N=20,000$ and $f(x,y)=\sin(x+y)$. The value of $\QC_N$ is indicated by color coding. The dark blue region is chaotic, and all other colors indicate quasiperiodicity. Convergence of $\Q_N$ in the quasiperiodic region is slower (yellow to green) when the rotation number of an orbit is close to a rational number $m/n$ where $n$ is small such as $1/5$ or $1/6$. \red{See Corollary 2.1 in \cite{Das-Yorke} for details of the calculation. When $N$ is increased to $10^6$, almost all of the quasiperiodic points in the $500\times500$-point set displayed become red. } }

\caption{\textbf{A quasiperiodic circle for the Standard Map with $r=1.0$.} Top left : The curve. Top right: The function $g(\theta) = \phi(\theta)-\theta$, the periodic part of the change of coordinates between the quasiperiodic circle and the pure rotation with rotation $\rho \approx 0.121$. Bottom left: The exponential decay of the Fourier coefficients of $g(\theta)$, only shown for even $k$ because $a_k=0$ for all odd $k$ . Note that $|a_{k}|=|a_{-k}|$. Bottom right: The super convergence of $\Q_N$ for the rotation number of Standard Map. \red{See Corollary 2.1 from }\cite{Das-Yorke} \red{for details of the calculation.} Note that the $|a_k|$ has a local maximum at $k=506$. \red{ %[[g]] This type of spike can occur in the presence of small divisors, but we have verified that the spike correctly reflects the Fourier series. See discussion.} }

\caption{\label{fig_Folded_Rot} \red{{\bf The folded curves ``the fish'' in Eq.~\ref{eqn:fish} and ``the flower'' in Eq.~\ref{eqn:flower}.} The curves wind $j$ times around $P_j$ .}}

312 — 1512.07598

\caption{(color online) Variation of the band structure of group IV-VI monolayers in the puckered configuration with small mechanical strains. Both compressive (negative) and tensile (positive) \textcolor{red}{strains} are applied along the lattice direction $a$ and $b$. The systems, mentioned above, undergo a indirect-to-direct band gap transition due to the application of very small strain ($\leq$ 3 $\%$). The values above sub-figure represent the amount of strain required to make this transition. The vertical arrows (blue color) indicate the vertical transitions from the local maxima of the valence band to the local minima of the conduction band. Numerical values written near the arrows represent the corresponding band gaps.}

313 — 1512.07877

\caption{Energy and enstrophy (scaled by $\alpha^2$) vs. time for the 3D Euler-Voigt equations. (red ``$+$'': $\|\bu^\alpha(t)\|_{L^2}^2$, blue ``\textasteriskcentered'': $\alpha^2\|\nabla\bu^\alpha(t)\|_{L^2}^2$, black ``$\circ$'': $\|\bu^\alpha(t)\|_{L^2}^2 +\alpha^2\ \nabla\bu^\alpha(t)\|_{L^2}^2$.) Resolution: $256^3$. }

314 — 1512.07908

\caption{ {$\mid$ \bf The critical exponent \boldsymbol{$\eta$}.} The critical exponent $\eta$ can be obtained from the fitting in the log-log plot of the Kekule-VBS structure factor versus $L$ at $J=J_c$ for different $N$. ({\bf a}) $N=2$. ({\bf b}) $N=3$. ({\bf c}) $N=4$. ({\bf d}) $N=5$. ({\bf e}) $N=6$. \\ }

\caption{{$\mid$ \bf Data collapse results for different \boldsymbol{$N$}.} The critical exponent $\nu$ can be obtained through data collapse. ({\bf a}) $N=2$. ({\bf b}) $N=4$. ({\bf c}) $N=5$. ({\bf d}) $N=6$. }

315 — 1512.08512

\caption{{\small (a) We measured the rate at which subjects chose an algorithm's synthesized sound over the actual sound. Our full system, which was pretrained from ImageNet and used example-based synthesis to generate a waveform, significantly outperformed models based on image matching. For the neural network models, we computed the auditory metrics for the sound features that were predicted by the network, rather than those of the inverted sounds or transferred exemplars. (b) What sounds like what, according to our algorithm? We applied a classifier trained on {\em real} sounds to the sounds produced by our algorithm, resulting in a confusion matrix (\cf Fig. \ref{fig:mean-coc}(b), which shows a confusion matrix for real sounds). It obtained \predsoundbalacc class-averaged accuracy. (c) Confusions made by a classifier trained on \fcseven features (\visualacc class-averaged accuracy). For both confusion matrices, we used the variation of our model that was trained from scratch (see Fig. \ref{fig:channel}(b) for the sound confusions made with pretraining).}}

316 — 1512.08582

\caption{The raw KLF for stars in the cluster region (cluster region KLF) and {for field objects} (field KLF) are shown by the gray solid line with gray-filled circles and gray dot-dashed line with gray-filled squares, respectively. % The star counts of field objects are normalized to match with the total area of the cluster region. % The KLF for the S207 cluster (S207 KLFs), shown by the black thick line with black-filled circles, is made by subtracting the {normalized} star counts in the field KLF from {star counts} in the cluster region KLF. % Error bars are the uncertainties from Poisson statistics. % The upper limit of the counts in $K=19$\,mag bin for the cluster region KLF is shown by the gray dashed line with the gray open circle considering the detection completeness, while that for the controld field KLF is shown by the gray dashed lines with the gray open square.% The resultant upper and lower limits of the counts in $K=19$\,mag bin for the S207 KLF are shown with the black dashed line with the black open circle and black dotted line with the black circle, respectively.% % The vertical dot-dashed line and vertical dashed line show the limiting magnitudes of the 10\,$\sigma$ detection (19.0\,mag) and 5\,$\sigma$ detection (20.0\,mag), respectively.}

317 — 1512.08689

\caption{\small Termination evaluation results. (a) Overview table. (b) Comparison of \tool{T2} and \tool{AProVE}. \textcolor{green}{Green} (resp. \textcolor{blue}{blue}) marks correspond to terminating (resp. non-terminating) examples, and \textcolor{gray}{gray} marks examples on which both provers failed. A $\square$ (resp. a $\triangle$) indicates an example in which only \tool{T2} (resp. \tool{AProVE}) succeeded, and $\circ$ indicates an example on which both provers return the same result.}

318 — 1512.08849

\caption{\textcolor{blue}{The results of testing on SICK data set by using the models trained only on SNLI data set.}}