2013

1 — 1301.0007

\caption{(a) The entire networks containing all nodes and all links in the identified $B$-motifs for four stocks: 000021 ({\color{red}{$\blacklozenge$}}), 000539 ({\small\color{magenta}{$\blacktriangle$}}), 200002 ({\tiny\color{green}{$\blacksquare$}}), 200625 ({\large\color{blue}{$\bullet$}}). Each edge in the plot represents a pair of bidirectional links. (b) The entire network containing all nodes and all links in the identified $C$-motifs for stocks 200024. Each edge stands for multiple directed links.}

2 — 1301.0111

\caption{a) Height to the Galactic plane of symmetry ($Z_{GC}$) as a function of the Galactic radius ($R_{GC}$) for the whole sample. Blue crosses represent stars from the \emph{LRc01} direction, green circles stars from the \emph{SRc01} direction, and red plus stars from the \emph{LRa01} direction. Planet-hosting stars are represented with black thick uncertainty bars. The dotted lines correspond to the typical scale height of the thin and thick discs reported by \citet{2008A&A...480..753V}. b) Toomre diagram representing the different Galactic populations in our sample: thin disc (cyan $\triangle$), thick disc (yellow \textasteriskcentered) and halo stars, (orange $\diamond$). Dotted lines represent circles with radius every 50~\kms. c) Bottlinger diagram: $U$ versus $V$ velocities. The symbols are the same as in the Toomre diagram. The thin disc stars are focussed at $U=0$~\kms~and $V=0$~\kms. d) Kinematical heat diagram of $W$ velocities. The symbols are the same as in the Toomre diagram. The separation between the stellar populations is clear with the kinematically hottest stars being the halo stars and the coolest ones corresponding to the thin disc. e) Peculiar velocity ($v_{pec}=\sqrt{U^{2}+V^{2}+W^{2}}$) as a function of the overall metallicity. The symbols are the same as in (f). We represented in thick black the planet-hosting stars for (a) to (e). f) $\alpha-$enhancement as a function of the metallicity. The solid line corresponds to the standard law we used in our grid, the dashed line corresponds to 1$\sigma$ deviation from this law.}

3 — 1301.1314

\caption{Potential gradient in thin-film \lvosto heterostructures. The curves (\textcolor{red}{\tiny$\blacksquare$}/% \textcolor{green}{\large\protect\raisebox{-.15ex}{\textbullet}}/% \textcolor{blue}{\footnotesize\protect\raisebox{.12ex}{$\blacktriangle$}} distinguish structure size) track the energy of an \Ox-1s state through the \ce{TiO2} ($\text{layer} \le 0$) and \ce{VO2} ($\text{layer} > 0$) layers, which provides a measure of the potential gradient. Above the data, the layers, interfaces (IF), and effective electron potential $V^*$ are shown schematically. The potential in the \sto part is almost flat, showing only a slight buckling which we attribute to band bending at the interface.}

\caption{Potential gradient in multi-layer \lvosto heterostructures. The curves (\textcolor{red}{\tiny$\blacksquare$}/\textcolor{green}{\large\protect\raisebox{-.15ex}{\textbullet}}/\textcolor{blue}{\footnotesize\protect\raisebox{.12ex}{$\blacktriangle$}} distinguish structure size) track the energy of an \Ox-1s state through the \ce{TiO2} ($\text{layer} \le 0$) and \ce{VO2} ($\text{layer} > 0$) layers, providing a measure of the potential gradient. Above the data, the layers, interfaces (IF), and effective electron potential $V^*$ are shown schematically. Periodicity forces the potential to return to zero in the \sto part after it has ramped up in the \lvo part.}

4 — 1301.1810

\caption{Temporal series of the impeller speeds $f_1$ (\textcolor{turq}{blue}) and $f_2$ (\textcolor{garnet}{red}) for various $\gamma$. $(a)$, symmetric high-speed state \st{s} observed at $\gamma = -0.0164$; $(b)$, threshold of the irregular oscillations \st{i_2} with very small events for $\gamma= -0.0460$; $(c)$, \st{i_2} irregular oscillations for $\gamma = -0.0668$; $(d)$, multi stable regime showing \st{s}, \st{i_2} and \st{b_2} events at $\gamma = -0.0891$; $(e)$, a single fast rare event in a quasi-steady slow \st{b_2} regime at $\gamma = -0.0912$; $(f)$, slow $(b_2)$ regime for $\gamma = -0.1049$.}

\caption{Joint-probability density maps of the $(f_1,f_2)$ values (log-scale), based on the temporal signals of fig.~\ref{fig:tempor}: $(a)$, steady symmetric high-speed state: $\gamma = -0.0164$; $(b)$, beginning of irregularly oscillating states towards an intermediate state, $\gamma = -0.0460$; $(c)$, typical irregular oscillations with large slowdowns: $\gamma = -0.0668$; $(d)$, multistability with three `most visited' states: $\gamma = -0.0891$, (\textcolor{blue}{---}) and (\textcolor{red}{---}) represent respectively the fig.~\ref{fig:mytransitions} mean profile for down and up transitions; $(e)$, rare events for $\gamma = -0.0912$; $(f)$, steady slow state: $\gamma = -0.1042$. The black dashed-dotted line represents the $\theta = 0$ condition. Both fast and slow states can be observed at $\theta = 0$ considering the shape of the fig.~\ref{fig:gammatheta} curve.}

\caption{Survival function $1 - \int p.d.f.$ of the distribution of the time spent in the slow quasi-steady state \st{b_1} before transiting to one of the faster quasi-steady states \st{i_1} and \st{s}. (\textcolor{teal}{$\bullet$}), experimental data, broken line, exponential fit with two characteristic times. Left, $\gamma = -0.094$ favours a simple exponential distribution. Right, $\gamma = -0.097$, closer to the edge of the steady $(b_2)$ branch, favours separate characteristic times.}

5 — 1301.2053

\caption{Bias (left) and misclassification rate (right) due to shift contamination for $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $h=\lfloor(n+p+1)/2\rfloor$ and Normal-distributed observations shown as a function of $\nu$. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines).}

\caption{Bias (left) and misclassification rate (right) due to shift contamination for $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $h=\lfloor(n+p+1)/2\rfloor$ and Cauchy-distributed observations shown as a function of $\nu$. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines).}

\caption{Bias (left) and misclassification rate (right) due to point-mass contamination for $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $h=\lfloor(n+p+1)/2\rfloor$ and Normal-distributed observations shown as a function of $\nu$. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines).}

\caption{Bias (left) and misclassification rate (right) due to point-mass contamination for $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $h=\lfloor(n+p+1)/2\rfloor$ and Cauchy-distributed observations shown as a function of $\nu$. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines).}

\caption{Bias (left) and misclassification rate (right) due to the Barrow wheel contamination for $\varepsilon=\{0.1,\ldots,0.4\}$ and $p=\{4,\ldots,16\}$, $h=\lfloor(n+p+1)/2\rfloor$. \textcolor{fmcd}{FastMCD}, \textcolor{fmve}{FastMVE}, \textcolor{msde}{SDE}, \textcolor{PCS}{FastPCS} (far right).}

\caption{Bias (left) and misclassification rate (right) due to the shift contamination for $h\approx 3n/4$, $\varepsilon=\{0.1,0.2\}$ and $p=\{4,\ldots,16\}$. The first (last) two rows are for multivariate normal (Cauchy) distributed r.v. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines).}

\caption{Bias (left) and misclassification rate (right) due to point-mass contamination for $h\approx 3n/4$, $\varepsilon=\{0.1,0.2\}$ and $p=\{4,\ldots,16\}$. The first (last) two rows are for multivariate normal (Cauchy) distributed r.v. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines). The online version of this figure is in color.}

\caption{Empirical bias at uncontaminated samples for $p=8$ as a function of sample size. \textcolor{fmcd}{FastMCD} (dotdash lines), \textcolor{fmve}{FastMVE} (longdash lines), \textcolor{msde}{SDE} (twodash lines), \textcolor{PCS}{FastPCS} (solid lines). The online version of this figure is in color.}

6 — 1301.2920

\caption{Size of entrapped bubbles $R_{bub}$, measured as a function of impact velocity $U$. The data summarize all collision events, for varying drop sizes. The triangles ($\triangleleft$) represent the data for floating bubbles. For these bubbles the colors indicate the approximate impact drop radius: red triangles (\textcolor{red}{$\triangleleft$}) $R<400$ $\mu m$, blue triangles (\textcolor{blue}{$\triangleleft$}) $400<R<600$ $\mu m$ and black triangles (\textcolor{black}{$\triangleleft$}) $R>600$ $\mu m$. The squares (\textcolor{blue}{$\square$}) denote the sticking bubbles for $0.21<R<0.62$ mm. The typical measurement error is indicated in the legend.}

7 — 1301.3769

\caption{{\color{blue}{Left: an example of AFM image of the surface of a Ag film. The flat right-hand side of the image is the glass substrate. The sharp edge shown here was obtained by scratching the film. The white straight line indicates the cut along with the z-profile shown in the right panel was measured. This profile indicates a thickness of $65 \pm7$ nm. The effective thickness of this film that results from the correction for voids as described in ref.\cite{Rowell03} and in the text is equal to 30 nm.}}}

\caption{Induced surface charge density $n_{2D}$ as function of the gate voltage $V_g$ for different devices made of different metals (Au, Ag, Cu). {\color{blue}{Although most of the points are gathered around a common trend, some data show much higher values of the surface charge density. Lines are only guides to the eye.}} In the inset the dependence of $n_{2D}$ on the sequence of application of $V_g$ is shown. For details see the text.}

8 — 1301.4802

\caption{\redc{The difference between the thermodynamic force of the potential interpolation-based scheme and force interpolation of the standard AdResS. The green solid line shows the difference of the thermodynamic forces, i.e. $\Delta{\vect F}_\thf={\vect F}_\thf^\hadress-{\vect F}_\thf$. The blue and red symbols indicate the quantity $\langle w\nabla w(U^{\AT}-U^{\CG})\rangle$ calculated in the two different approaches (blue corresponds to the standard AdResS while red to the Hamiltonian-based approach). The equivalence between the two approaches is suggested by the fact that $\Delta{\vect F}_\thf$ corresponds to $\langle w\nabla w(U^{\AT}-U^{\CG})\rangle$, and that this latter is actually the same independently from the approach used to calculate it.} }

9 — 1301.4861

\caption{The first twenty singular values of the balanced ARE solutions for sector balanced truncation for sector $[-2830,~14200]$ (\textcolor{blue}{*}) and the Hankel singular values (\textcolor{red}{+}).}

\caption{$Re=2000$, $k=0$, $n=1$. The upper panel shows the eigenvalues of Eq.~(\ref{matrixsectordefn}) for the original model (-), the sector-balanced reduced model (\textcolor{blue}{-$\cdot$-}) and for balanced truncation (\textcolor{red}{-~-}). The balanced truncation method violates the inequality of Eq.~(\ref{matrixsectordefn}) showing that the reduced model is no longer in the sector chosen. The lower panel shows the fractional error of the reduced models compared with the original.}

10 — 1301.4968

\caption{\textbf{Distributional effect on accuracy}: parameter-estimate distributions for normal $\boldsymbol{\mathcal{N}}$ and $\boldsymbol{\chi^2}$ noise, both having Gaussian-shaped autocorrelation functions with correlation time $\tau$, a sampling frequency of $1/2\tau$ and a sample period of $32 \, \tau$. The smoothed histogram of 1000 $\tau$ estimates are compared for maximum-likelihood \textcolor{black}{$\boldsymbol{-}$} and variogram regression \textcolor{red}{$\boldsymbol{\cdots}$}. The accuracy advantage in the likelihood estimator is diminished, but still present, as the distributional assumptions are violated.}

\caption{\textbf{Sample-size effect on accuracy}: parameter-estimate distributions for Gaussian-shaped autocorrelation functions with correlation time $\tau$. The \textbf{M}edium sample has a sampling frequency of $1/2\tau$ and a sample period of $32 \, \tau$, while the \textbf{S}mall sample has a sampling frequency of $1/\tau$ and a sample period of $16 \, \tau$. The smoothed histogram of 1000 $\tau$ estimates are compared for maximum-likelihood \textcolor{black}{$\boldsymbol{-}$} and variogram regression \textcolor{red}{$\boldsymbol{\cdots}$}. The likelihood estimate undergoes rapid improvement with sample size as compared to the regression estimate.}

\caption{\textbf{Continuity effect on accuracy}: parameter-estimate distributions for model \eqref{eq:rational} with correlation time $\tau$ and varying degree $d=0,2$ of continuity $C^d$, given a sampling frequency of $1/2\tau$ and a sample period of $32 \, \tau$. The smoothed histogram of 1000 $\tau$ estimates are compared for maximum-likelihood \textcolor{black}{$\boldsymbol{-}$} and variogram regression \textcolor{red}{$\boldsymbol{\cdots}$}. In terms of sample size, it is easier to achieve a marked improvement in parameter estimation with continuous processes.}

\caption{\textbf{Reliability of confidence intervals}: The distribution of (squared) parameter errors relative to the estimated errors, from the \textbf{M}edium sample in Figure~\ref{fig:size}, for maximum likelihood $\boldsymbol{-}$ and variogram regression \textcolor{red}{$\boldsymbol{\cdots}$}. The variogram distribution here is extremely long tailed even though its parameter distribution is not. The regression estimate suffers from a severe underestimation of standard errors. }

\caption{\textbf{Limiting behavior of variogram-derived standard errors}: from the \textbf{S}mall \textcolor{yellow}{$\boldsymbol{-}$} and \textbf{M}edium \textcolor{orange}{\textbf{- -}} samples of Figure~\ref{fig:size}, and a larger sample \textcolor{red}{$\boldsymbol{\cdots}$}, with further doubled period and frequency. The parameter-estimate distributions are converging with increasing sample size, albeit much more slowly than with ML, however the performance of the standard error estimate becomes increasingly poor.}

11 — 1301.5546

\caption{{\it IV} characteristics of sample \#1 at a few values of Josephson coupling energy:$E_J$ = 9 $\mu$eV ({\color{black} black}), 7.3 $\mu$eV ({\color{red}red}), 6 $\mu$eV ({\color{blue}blue}), 4.5 $\mu$eV ({\color{magenta}magenta}), and 2.8 $\mu$eV ({\color{OliveGreen}green}), without base current ($I_B=0$) at $T=90$ mK.}

\caption{(a) Middle traces: {\color{magenta}magenta} and black IV curves are measured without base current at $E_J = 6.5$ and 5.8 $\mu$eV, respectively. {\color{red}Red curves}, corresponding to $E_J = 6.5$ $\mu$eV are measured at $I_B=+0.3$, +0.34, and +0.38 nA (traces from right to left). {\color{blue}Blue curves} have the same bias conditions as the red curves but were measured at $E_J = 5.8$ $\mu$eV. The red curves are offset by (+0.22 mV, +0.42 nA) for clarity, like the blue curves by (+0.22 mV, -0.42 nA).\\ (b) The normal operation region of the BOT at $E_J=7.1$ $\mu$eV with increasing $I_B$. Negative slope is the Landau-Zener tunneling regime, increases with $I_B$ and eventually the slope diverges: $I_B = +0.06$, 0.065, 0.07, 0.075, 0.08, 0.085, 0.095, and 0.105 nA (from right to left). Filled (open) circle traces are of $I_E$ when $V_C$ is swept from left (right) to right (left).The measurement temperature was at $T\sim 90$ mK. }

12 — 1301.5971

\caption{Mass of DM \textcolor{red}{minispike} within orbital radius $r$. A steeper density distribution contains more DM mass within the radius $r$. The solid line is $\alpha = 1.5$, the dashed line is $\alpha = 2.0$, and the dot-dashed line is $\alpha = 7/3$.}

13 — 1301.6341

\caption{\small Intersection of two one-way streets of width $M$. The blue particles (\bluew{$\blacktriangleright$}) move eastward and the orange particles (\orangew{$\blacktriangle$}) northward. The parameter $\alpha$ determines the particle injection rate. The region bordered by the heavy solid line is the `intersection square'. Figure taken from \cite{cividini_a_h2013a}. }

\caption{\small Intersection of two single lanes. (a) A waiting line of length $\ell(t)$ at the intersection site divides the horizontal lane into a free flow and a jammed flow domain.\,\, (b) The random phases of the particles in the horizontal lane define a division of them into platoons. A heavy vertical red bar has been placed to the left of the last particle of each platoon. Those in the jammed flow domain are compact.}

14 — 1301.6662

\caption{A singular extremal trajectory composed of a $\phi_\omega$-singular primitive, plotted in \textcolor{red}{red}, and regular primitives, plotted in \textcolor{blue}{blue}. In this case, the $\phi_\omega$-singular primitive does not contain a straight line segment. The line $\ell$ is defined by $c_1 y - c_2 x + c_3 = 0$. }

\caption{A singular extremal trajectory composed of a $\phi_\omega$-singular primitive, plotted in \textcolor{red}{red}, and regular primitives, plotted in \textcolor{blue}{blue}. In this case, the $\phi_\omega$-singular primitive contains a straight line segment. The line $\ell$ is defined by $c_1 y - c_2 x + c_3 = 0$. }

15 — 1301.7435

\caption{Two topological equivalent pumping lattices. (a). A sliding lattice ($V_{1}=0,V_{2}=1E_{R}$) (b). A continuous Rice-Mele pump \cite{Rice:1982p61164} ($V_{1}=2E_{R},V_{2}=1E_{R}$), where the dimerization of hopping amplitudes and onsite energies are modulated periodically. At $t=0$ and $t=T/2$ different topological phases of the Su-Schrieffer-Heeger lattice \cite{Su:1979p972} are realized. (c). A tight-binding schematic view the pumping process \red{in} (b). The two pumping processes (a) and (b) are topologically equivalent for the lowest band and can be adiabatically connected, see the main text. }

\caption{(a). Spatial-temporal structure of the optical lattice Eq. (\ref{eq:lattice}) for $V_{1}=V_{2}=4E_{R}$. (b). Berry curvature distribution of the lowest band of $H(k_{x},t)$. Integration over the Brillouin zone shows the Chern number equals to \red{$1$}. (c). Pumped charge and total current Eq.~(\ref{eq:pumpedcharge}) of an infinite sized system with $T=40\hbar/E_{R}$ and one particle per unit-cell. The pumped charge is quantized at full pumping cycles. }

16 — 1301.7515

\caption{System Parameters Used in Numerical Evaluation.}{\includegraphics[scale=0.7]{Tab1.eps}\label{Tab1}}

17 — 1301.7536

\caption{The branching ratios of $\tilde a$ into $\gamma \nu$ ({\color{red} red}), $Z \nu$ ({\color{green} green}) and $W l$ ({\color{blue} blue}) as a function of $m_{\tilde B}/m_{\tilde W}$ for $3C_Y/5 = C_W$ (left) and as a function of $5C_W/3C_Y$ for $m_{\tilde B}/m_{\tilde W} = 0.5$ (right). We assume $m_{\tilde a} \simeq 260\,$GeV and $|m_{\tilde B} - \mu|, |m_{\tilde W} - \mu| \gg m_Z$.}

\caption{(Top) Constraints on the plane of $\kappa$ and $f_a$ for the bino MSSM-LSP with $C_W=0$, $C_Y=1$, $m_{\tilde a}=260$\,GeV,$m_{\tilde B}=1$\,TeV,$m_\sigma=500$\,GeV and$\sigma_i = f_a$. In the light blue band, the Fermi 130\,GeV gamma-ray line excess is explained. On the right side of the black dashed line, the bino lifetime is shorter than$0.1$\,sec. Above the orange dot-dashed line, the axino abundance produced by the bino decay satisfies$\Omega_{\tilde a}^{(\tilde B)} h^2 < 0.1$, where $\Omega_{\tilde B}h^2 = 10$ is taken as a reference. For each $f_a$, $T_{\rm R}$ is set so that the axino becomes the dominant component of the DM. Below the red horizontal line, $\Delta N_{\rm eff} < 1$ is obtained from the axions produced by the saxon decay. Above the blue dotted horizontal line, $T_{\rm R} > m_{\tilde a}$ is satisfied to account for the axino DM. (Bottom) Same as the top panel, but for the wino MSSM-LSP with $C_W=1$, $C_Y=0$ and $m_{\tilde W}=1$\,TeV.}

18 — 1302.0021

\caption{Experimental data for Ne ionization yields measured at 93.0 eV (\fullsquare) and 90.5 eV (\fulltriangledown) photon energy with pulses of 15 fs duration compared with results from the present minimal model (\full) and results from the elaborate rate description for 30 fs pulse length of Ref. \cite{Lambropoulos2} (\dashed). \label{fig3}}

19 — 1302.0281

\caption{(Color online) Magneto-association induced loss at $B = 734.5\,$G, where $a \simeq 1100\,a_0$. The main plot shows loss spectra for thermal gases with the following modulation amplitudes and approximate temperatures: ({\color{OliveGreen} $\blacktriangle$})~0.57\,G\,/\,$10\,\uK$, ({\color{RoyalBlue} $\blacksquare$})~0.14\,G\,/\,$3\,\uK$, and ({\color{Red} \large$\bullet$})~0.57\,G\,/\,$3\,\uK$. The solid curves are fits to Lorentzians convolved with thermal Boltzmann distributions. For the 0.57\,G\,/\,$3\,\uK$ data the lower frequency resonance is a subharmonic response, while the primary resonance is thermally broadened by strong modulation. The inset ({\color{Black} $\blacklozenge$}) corresponds to a BEC with a modulation amplitude of 0.14\,G. The solid black line is a Lorentzian fit to the condensate resonance and the vertical dashed line in the main figure is the resonance location$E_\mathrm{b} / h = 450\,$kHz found from this fit.}

\caption{(Color online) Results of modulation spectroscopy using condensates ({\color{Red} $\bullet$}) and $\sim$$3\,\uK$ thermal clouds ({\color{Black} $\blacksquare$}). (a) and (b): $E_\mathrm{b}$ vs.~$B$, and (c) and (d): the same data plotted as $\gamma \equiv (mE_\mathrm{b}/\hbar^2)^{1/2}$ vs.~$B$. The vertical error bars correspond to uncertainty in fitting to the binding energy resonances, while the horizontal error bars are the statistical uncertainties due to shot-to-shot variations of the magnetic field. The relatively large error bars below 725\,G arise from the broadening of the resonance from the strong modulation required to produce a detectable signal. The lines are fits of the measurements to Eq.~(\ref{eqn:fr}) using the various models. Solid (black): universal model, $E_\mathrm{b} = \hbar^2/ma^2$; dashed (green): simple two-channel model, Eq.~(\ref{eqn:binding}), with the parameters given in the text; dot-dashed (blue): complex two-channel model given in Ref.~\cite{Prico11} (Eq.~26), again with parameters as given in the text (the two-channel model of Ref.~\cite{Lange09} gives similar results); and dotted (red): Eq.~(\ref{eqn:gamma}) using $R_\mathrm{e}$($B$) from Fig.~\ref{fig:FB_RE}. The fits exclude data below 725\,G, where$a \simeq 250~a_0$ is no longer much greater than $R^*$, and the validity of the non-universal corrections becomes questionable. While the nominal fitting parameters are $a_\mathrm{bg}$, $\Delta$, and $B_\infty$, $\Delta$ is fixed at -174\,G. The fits are weighted by the inverse uncertainties. The resulting Feshbach resonance parameters are given in Table~\ref{table:parameters}. }

20 — 1302.1208

\caption{List of VHE-\gray~spectra included in the analysis. The table shows the redshift of the source, the IACT experiment that measured it, the energy range covered by the spectrum and the number of data points in the optical thick regime for the optical depth given by the \KD~and by the scaled version of the \FRV. }

\caption{Significance map for the \pa~conversion in the $(\ma,\gag)$ plane. Smaller values (brighter regions) indicate less accordance between the model and the data. Upper panel: $p_t$ values for the \scenmax- scenario, shown as $-\log_{10}(p_t)$. Lower panel: $p_{95}$ values for the \scenICM~case. For each pixel, 5000 realizations of the random magnetic field are simulated and $p_{95}$ is determined from the resulting 5000 $p_t$ values (cf. Sec. \ref{sec:method}). In the left column, the attenuation due to the interaction of VHE \gray s with the EBL is given by the \KD, whereas in the right column the \FRV~is utilized. The maps are smoothed using a bilinear interpolation between the single $(\ma,\gag)$ pixels.}

\caption{List of fit qualities for all VHE-\gray~spectra if no ALPs are included in the de-absorption of the spectra. The table shows the $\chi^2$ values, the degrees of freedom (d.o.f.), and the resulting fit probabilities $p_\mathrm{fit}$ for both EBL models used here. See Table \ref{tab:spectra} for the references of each spectrum. }

21 — 1302.1673

\caption{\label{fig3} BDL dwell (\textcolor{blue}{$\triangle$}) and closure times (\textcolor{blue}{$\circ$}) ($\beta\kappa_\phi=300$), and ADL dwell times (\textcolor{red}{$\Diamond$}) ($\beta\kappa_\phi=200$) with fits (see text). Horizontal arrows are the TA dwell times for 3 values of $\beta\kappa_\phi$.}

22 — 1302.2065

\caption{An energy-redshift plot of the \gray horizon showing our uncertainty band results \protect\cite{sms12} compared with the {\it Fermi} plot of their highest energy photons from FSRQs (red), BL Lacs (black) and GRBs (blue) vs. redshift \protect\cite{abd10}.}

23 — 1302.2511

\caption{\label{figrdist} Continuous size distributions used to create the soft sphere packings. The distributions are: (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_A = 0.40$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_A = 0.44$; (\textcolor{Purple}{$\square$}) uniform $\sigma_A = 0.44$. The open symbols represent the original size distribution and the closed symbols represent the size distribution once rattlers are removed.}

\caption{\label{figphiRCP} (a) $\phi_{RCP}$ versus the standard deviation of the surface area distribution $\sigma_A$ for a variety of size distributions in three dimensions: (\textcolor{OliveGreen}{$\triangledown$}) monodisperse; (\textcolor{blue}{$\lozenge$}) bidisperse; (\textcolor{Purple}{$\square$}) uniform; (\textcolor{brown}{$\vartriangle$}) Gaussian; (\textcolor{black}{$\bigcirc$}) lognormal. $\phi_{RCP}$ including all particles is plotted with open symbols while solid symbols correspond to $\phi_{RCP}$ with rattlers omitted. The dashed line indicates the average $\phi_{RCP}$ with rattlers omitted for all size distributions. Inset: Percentage of rattlers at $\phi_{RCP}$ in three dimensions versus the standard deviation of the surface area distribution $\sigma_A$. (b) $\phi_{RCP}$ versus the standard deviation of the radius distribution in two dimensions with (open symbols) and without rattlers (closed symbols). The dashed line is the average $\phi_{RCP}$ with rattlers omitted for all size distributions with $\sigma_R \geq 0.1$. Inset: Percentage of rattlers at $\phi_{RCP}$ in two dimensions and the standard deviation of the radius distribution $\sigma_R$. The symbols correspond to the same data as in Figure \ref{figphiRCP}(a). }

\caption{\label{figlewis} The average of the contact number distribution for a particle of a given size for the six different size distributions at $\phi_{RCP}$: (\textcolor{OliveGreen}{$\triangledown$}) monodisperse; (\textcolor{blue}{$\lozenge$}) bidisperse, radius ratio 1:1.4; (\textcolor{Purple}{$\square$}) uniform $\sigma_A = 0.23$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_A = 0.27$; (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_A = 0.40$; (\textcolor{red}{$\vartriangleleft$}) lognormal $\sigma_A = 0.61$; (\textcolor{cyan}{$\ast$}) lognormal $\sigma_A = 0.72$. We present three different scalings: (a) in terms of the normalised radius $r$; (b) in terms of the normalised area $a$; (c) in terms of the normalised volume $v$. The data are plotted over a range that illustrates the quality of the collapse.}

\caption{\label{figlewis_a} The average contact number for particles of a given area $a$ at $\phi_{RCP}$ in three dimensions for all (\textcolor{OliveGreen}{$+$}) monodisperse, (\textcolor{blue}{$\lozenge$}) bidisperse, (\textcolor{Purple}{$\square$}) uniform, (\textcolor{brown}{$\vartriangle$}) Gaussian and (\textcolor{black}{$\bigcirc$}) lognormal size distributions at all the widths $\sigma_A$ we have considered (see Figure \ref{figmu2_vr}). The solid red line is a fit to Equation (\ref{eqzafit}). Inset: the average contact number for particles of a given radius $r$ in two dimensions at $\phi_{RCP}$. The solid red line is a fit to Equation (\ref{eqzrfit}). }

\caption{\label{figPzapred} The contact number distribution for particles of a given size $P(z|a)$ for four size distributions at $\phi_{RCP}$ in three dimensions: (\textcolor{OliveGreen}{$\triangledown$}) monodisperse; (\textcolor{Purple}{$\square$}) uniform $\sigma_A = 0.44$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_A = 0.44$; (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_A = 0.40$. Plotted here is a selection of the $P(z|a)$ for 6 different intervals $a$ with the $P(z|a)$ shifted for clarity. Plotted in order of lowest to highest is $0.475< a < 0.525$; $0.975 < a < 1.025$; $1.475 < a < 1.525$; $1.975 < a < 2.025$; $2.475 < a < 2.525$; $2.975 < a < 3.025$. The solid red line is the model prediction of $P(z|a)$ from Equation (\ref{eqMOD}).}

\caption{\label{figlewis_pred} (a) The average of the contact number distribution for particle of a given size for five different size distributions at $\phi_{RCP}$. The model prediction from Equation (\ref{eqzapred}) is plotted as the solid red line. (b) The variance of the contact number distribution for particle of a given $a$. The model prediction from Equation (\ref{eqmupred}) is plotted as the solid red line. (c) The ratio of the variance to the average of the contact number distribution for particles of a given size $a$. The dashed line denotes the value of the acceptance probability $p$ as found from the data. The size distributions and symbols plotted in all three panels are the same as in Figure \ref{figPzapred} with the addition of: (\textcolor{blue}{$\lozenge$}) bidisperse, radius ratio 1:1.4 and (\textcolor{cyan}{$\ast$}) lognormal $\sigma_A = 0.72$.}

\caption{\label{figlewis2D} The average of the contact number distribution for a particle of a given size for the six different size distributions in two dimensions at $\phi_{RCP}$: (\textcolor{blue}{$\lozenge$}) bidisperse, radius ratio 1:1.4; (\textcolor{Purple}{$\square$}) uniform $\sigma_R = 0.17$; (\textcolor{green}{$\vartriangleleft$}) lognormal $\sigma_R = 0.10$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_R = 0.24$; (\textcolor{magenta}{$\triangledown$}) lognormal $\sigma_R = 0.35$; (\textcolor{cyan}{$\triangleright$}) lognormal $\sigma_R = 0.45$. We present three different scalings: (a) in terms of the normalised radius $r$; (b) in terms of the normalised area $a$; (c) in terms of the normalised volume $v$. The data are plotted over a range that illustrates the quality of the collapse.}

\caption{\label{figzr}(a) The average of the contact number distribution for particles of a given radius $r$ in two dimensions for six different size distributions at $\phi_{RCP}$:(\textcolor{blue}{$\lozenge$}) bidisperse, radius ratio 1:1.4; (\textcolor{Purple}{$\square$}) uniform $\sigma_R = 0.17$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_R = 0.24$; (\textcolor{green}{$\vartriangleleft$}) lognormal $\sigma_R = 0.20$; (\textcolor{magenta}{$\triangleright$}) lognormal $\sigma_R = 0.30$; (\textcolor{cyan}{$\bigcirc$}) lognormal $\sigma_R = 0.40$. The model prediction is plotted as the solid red line. (b) The variance $\avg{ \sigma_{Z}^{2}|r }$ for the same size distributions. The model prediction is plotted as the solid red line. (c) The ratio of the variance to the average of P$(z|r)$. The dashed red line denotes the value of the acceptance probability $p$ as found from the data.}

\caption{\label{figaz}The average area of particles with a given contact number for four different size distributions in three dimensions at $\phi_{RCP}$: (\textcolor{magenta}{$\vartriangleleft$}) lognormal $\sigma_A = 0.10$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_A = 0.27$; (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_A = 0.40$; (\textcolor{Maroon}{$\vartriangleright$}) lognormal $\sigma_A = 0.82$. The solid lines are fits to Equation \ref{eqaz}. Top inset shows the average area of particles with a given contact number for all (\textcolor{black}{$\bigcirc$}) lognormal and (\textcolor{brown}{$\vartriangle$}) Gaussian size distributions at $\phi_{RCP}$, collapsed after rescaling by fitting the data to Equation (\ref{eqaz}). The dashed red line corresponds to a slope of 1. Inset of the top inset shows the fit parameter $\lambda$ as a function of $\sigma_A$. Bottom inset shows relationship between $\avg{a|z}$ versus $z$ for the same size distributions as plotted in Figure \ref{figlewis_pred} using the same symbols.}

\caption{\label{figrz}The average radius of particles with a given contact number for four different size distributions in two dimensions at $\phi_{RCP}$: (\textcolor{magenta}{$\vartriangleleft$}) lognormal $\sigma_R = 0.10$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_R = 0.24$; (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_R = 0.30$; (\textcolor{Maroon}{$\vartriangleright$}) lognormal $\sigma_R = 0.45$. The solid lines are fits to Equation \ref{eqrz}. Top inset shows the average area of particles with a given contact number for all (\textcolor{black}{$\bigcirc$}) lognormal and (\textcolor{brown}{$\vartriangle$}) Gaussian size distributions at $\phi_{RCP}$, collapsed after rescaling by fitting the data to Equation (\ref{eqrz}). The dashed red line corresponds to a slope of 1. Inset of the top inset shows the fit parameter $\lambda$ as a function of $\sigma_A$. Bottom inset shows relationship between $\avg{a|z}$ versus $z$ for the size distributions: (\textcolor{blue}{$\lozenge$}) bidisperse; (\textcolor{Purple}{$\square$}) uniform $\sigma_R = 0.23$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_R = 0.27$; (\textcolor{black}{$\bigcirc$}) lognormal $\sigma_R = 0.35$.}

\caption{\label{figAlaw}(a) Correlations between size of spheres in contact. All symbols represent the same size distributions as in Figure \ref{figAW5} except: (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_A = 0.27$. The solid lines are fits to the data using Equation (\ref{eqAlaw}) with the dashed lines being the uncorrelated prediction calculated from Equation (\ref{eqauncz}). The inset shows the fit parameter $w$ as a function of $\sigma_A$. The data in the inset is labelled the same as in Figure \ref{figphiRCP}(a). (b) Correlations between size of discs in contact in two dimensions rescaled by the predicted uncorrelated radius $\overline{R}_{nn}$. Five different size distributions are plotted; (\textcolor{blue}{$\lozenge$}) bidisperse, radius ratio 1:1.4; (\textcolor{Purple}{$\square$}) uniform $\sigma_R = 0.17$; (\textcolor{brown}{$\vartriangle$}) Gaussian $\sigma_R = 0.14$; (\textcolor{green}{$\vartriangleleft$}) lognormal $\sigma_R = 0.20$; (\textcolor{magenta}{$\triangledown$}) lognormal $\sigma_R = 0.30$; (\textcolor{cyan}{$\triangleright$}) lognormal $\sigma_R = 0.40$}

24 — 1302.2749

\caption{ECDFs of sojourn times for the \texttt{FB2010} dataset. The \map{} phase (left) shows significant improvements of HFSP over FAIR. Shorter sojourn times in the \map{} phase are reflected in the \reduce{} phase (middle), which shows that HFSP outperforms FAIR in terms of sojourn times. The ECDF of aggregate job sojourn times (right) indicates that both median and worst-case sojourn times are better with HFSP. The vertical line at 60 minutes (the workload has been consumed) indicates 20\% of completed jobs with HFSP vs. 1\% of completed jobs with FAIR.}

\caption{Swap times for a small cluster, using a synthetic workload of two jobs with one \reduce{} task each. The second \reduce{} task has a variable memory footprint, which corresponds to the amount of data swapped to disk by the OS, as reported on the x-axis.}

25 — 1302.2926

\caption{The \siline\feature of SN~2012fr (thick blue line) at +8 days in velocity space centered at the velocity minimum. Plotted for comparison are SN~2005hj\citep{quimby07} at +9 days in red, SN~1994D at +11 days \citep[from][]{blondin12} in magenta, and SN~2002bo \citep{benetti04} at +5 days in green, in order of thickest to thinnest lines.}

\caption{Left: Evolution of the \siline\line of SN~2012fr in velocity space. Right: Several representative examples of the two-component Gaussian fits. Data are in blue, regions of the pseudo-continuum fit are denoted by the vertical dotted black lines, the fitted pseudo-continuum is the dashed green line, the fully fitted profile is shown as the smooth red line, and the two individual components are the dotted cyan curves.}

\caption{Velocity evolution of the \siline\feature of SN~2012fr compared to that of a number of\sneia\from the\citet{benetti05} sample. The HVF component of SN~2012fr is shown as the large black triangles, while the lower velocity photospheric component is shown as large black circles.}

\caption{Pseudo-equivalent width (pEW) of the \siline\components in SN~2012fr as a function of phase. When both the HVF and photospheric components are clearly detected, they are shown as open blue triangles and open green circles, respectively. The total pEW of the\siline\line is shown as filled red squares. For references, the BSNIP sample\citep{bsnip1, bsnip2} is show as small black points.}

\caption{Late-epoch evolution of the \canir\in SN~2012fr, showing the narrow velocity width and constant velocity of the three lines comprising the triplet. For reference, we mark the wavelengths of the three triplet lines at a velocity of$v =$ 12,000~km~s$^{-1}$, showing that the reddest line of the triplet at 8662~\AA\is cleanly resolved from the bluer two lines.}

\caption{Velocity evolution of the high-velocity (triangles) and photospheric (circles) components of the \canir\(open red symbols) in SN~2012fr compared to that of the \siline\line (filled blue symbols).}

\caption{Fit to the \cahk\line profile in the maximum-light (\wifes\Nov. 11.67) spectrum of SN~2012fr. Line colors and styles are the same as in Figure~\ref{fig:twopeak_si_fits}, but with the \siblue\line profile plotted as the dashed-dotted black line. The top panel shows a fit without the\siblue\line, while the bottom panel shows a fit including this line.}

\caption{Spectral indicators from SN~2012fr compared to those of other \sneia\as measured by\citet{blondin12} and \citet{bsnip2}. Top Left: pEW(5972) vs. pEW(6355) at maximum light with spectroscopic subclasses as defined by \citet{branch09}. Top Right: pEW(6355) vs. $v_{\rm Si}$ at maximum light with velocity-based subclasses as defined by \citet{wang09}. Bottom Left: Velocity gradient of the \siline\line$\dot{v}_{\rm Si}$, as measured from the absolute decline between phases 0 and +10 days, vs. light-curve decline $\Delta m_{15}(B)$, as previously inspected by \citet{benetti05}. Bottom Right: Velocity gradient vs. $v_{\rm Si}$ at maximum light.}

26 — 1302.3022

\caption{Information on each test subject. \textasteriskcentered Number of days elapsed after death prior to testing \label{details}}

27 — 1302.3820

\caption{The mean-removed RSS signal $y_l[n]$ for four links that best show breathing during `sofa' experiment, using (top) basic and (bottom) breakpoint methods, with estimated breakpoints (\textcolor{darkgreen}{\bf :}).}

\caption{Normalized average PSD from (\ref{E:PSD}) using basic and breakpoint methods and data from Figure \ref{F:RSSSignal_2}, estimated breathing rate ($\bullet$ and $\square$), and true breathing rate (\textcolor{darkgreen}{\bf :}).}

\caption{Breakpoint method nap experiment breathing rate estimates based on 30 s data window (\textcolor{red}{$\cdot$}), 90 s median (\textcolor{Mulberry}{-----}), and breakpoints (\textcolor{green}{$\vdots$}) which indicate periods when movement is detected. }

\caption{Basic method nap experiment breathing rate estimates based on 30 s data window (\textcolor{red}{$\cdot$}), and 90 s median (\textcolor{Mulberry}{-----}). }

\caption{(a) lines for the top ten links by $v_l$, and (d) breathing image $\hat{x}$ (white=high), with true ($\boldsymbol{\times}$) and estimated (\textcolor{red}{$\boldsymbol{+}$}) locations.}

28 — 1302.4066

\caption{(Color online) The vectors $\mathbf{y}_i$ (\textcolor{red}{$\blacksquare$}) and $\mathbf{z}_i$ (\textcolor{black}{$\circ$}) of the response R\"ossler system~(\ref{eq:Roesslers}) for $\varepsilon=0.3$. The numbers $i$ of the vectors are shown by the regular and italic fonts, respectively}

\caption{(Color online) The vectors $\mathbf{y}_i$ (\textcolor{red}{$\blacksquare$}) and $\mathbf{z}_i$ (\textcolor{black}{$\circ$}) of the response R\"ossler system~(\ref{eq:Roesslers}) for $\varepsilon=0.3$, the length of the phase tube is $\tau=100$. The numbers $i$ of the vectors are shown by the regular and italic fonts, respectively}

\caption{(Color online) The vectors $\mathbf{y}_i$ (\textcolor{red}{$\blacksquare$}) and $\mathbf{z}_i$ (\textcolor{black}{$\circ$}) of the second generator with tunnel diode~(\ref{eq:KPRs}) for $\varepsilon=0.15$. The numbers $i$ of the vectors are shown by the regular and italic fonts, respectively. (\textit{a}) All neighbor vectors are used; (\textit{b}) only vectors whose phase trajectories pass through the phase tube with length $\tau=110$ are used}

29 — 1302.4534

\caption{(color online) Lattice dielectric constants ($\kappa_L$) for the amorphous structures of Hf$_{1-x}$Zr$_x$O$_2$. All the black symbols are for the values in each direction of $X$, $Y$ and $Z$ averaged over $10$ samples, while the square symbols connected with a solid red line represent the values which are averaged over all the three directions for each value of $x$. \textcolor[rgb]{0.00,0.00,0.00}{In the inset the standard deviations of the dielectric constant values over the $10$ samples along the $X$ direction have been presented.} }

\caption{Calculated lattice dielectric constants for HfO$_2$ and ZrO$_2$ \textcolor[rgb]{0.00,0.00,0.00}{relative to the dielectric constant at vacuum.}}

30 — 1302.4764

\caption{ The flux error due to truncation/aliasing of wave numbers with $q > \pi$ as a function of PSF size in pixels for a double Gaussian (lower panel), with the absolute value shown in the upper panel. The flux error is $1 - $measured/true, where `measured' refers to our $\sinc$-integration algorithm (aperture radius of 6 pixels), and `true' refers to the exact integral of the double Gaussian (sum of circular bivariate Gaussians having different amplitudes{\highlight (wing being 10\% of core)} and widths ($\sigma_{\mathrm{wing}}=2\sigma_{\mathrm{core}}$)). Curves are shown for a double Gaussian centred on an integer pixel value (solid), and one shifted by 0.5 pixels (dashed). Limits based on our conservative theoretical estimate are shown with dot-dashed lines. The theoretical estimate is the limit for truncated flux, while the $\sinc$-integrated curves include aliased flux. Dotted lines indicate a fractional error of 0.001, or $\sim$1 mmag. For the Gaussian PSFs, the flux error is $\ll 1$mmag provided sampling is $\sigma \gtrsim 1$ pixel.{\highlight In cases of severe undersampling, the formalism of including the pixel response in the double-Gaussian PSF breaks down; therefore, values for small PSF widths ($\sigma_{\mathrm{core}} < 0.4$ pixels) are not shown.} }

\caption{\highlight Measured growth curves and residuals with respect to computed theoretical fluxes. Left and right panels show values for double-Gaussian and Moffat PSFs, respectively. PSF widths of $\sigma$=0.8, 1.0, and 1.2 pixels were used to evaluate under-, critical-, and over-sampling. The differences between the measured and theoretical curves are too small to be seen directly, but are shown as residuals in the three lower panels. Residuals shown with a dashed line are the minimum observed, and correspond to no sub-pixel shift. Those shown with a solid line are the maxima, and correspond to a half-pixel shift in both $x$ and $y$ coordinates. The curves are very similar in structure, but differ dramatically in amplitude. The method is clearly unsuitable for sub-sampled ($\sigma < 1.0$ pixel) PSFs, but performs extremely well for $\sigma \gtrsim 1.2$ pixel.}

\caption{{\highlight Sinc aperture photometry performed on a simulated LSST $i'$ image \citep[][in preparation]{peterson13}. For clarity, only 1/4 of the values from a full frame are shown. The upper panel shows $m_\mathrm{ap}-m_\mathrm{cat}$ versus $m_\mathrm{cat}$, and the lower panel shows the standard deviation of the upper panel data in 0.5 mag bins. The aperture radius was 7 pixels (1.4\arcsec), and magnitudes are compared to the exact known values (i.e. $m_\mathrm{cat}$) used as input for the simulator.} }

31 — 1302.7186

\caption{\label{Fig:Exp} Saturation wavelength of the ripple instability induced in PVA films subject to flow. The wavelength does not vary with the flow speed or film viscosity. (a) All data. PVA ($\blacksquare$) $\eta=124$\,Pa\,s, ($\medbullet$) $\eta=25$\,Pa\,s and ($\blacktriangle$) $\eta=398$\,Pa\,s at ($\blacksquare$) 0.12\,m/s, (\textcolor[gray]{0.5}{$\blacksquare$}) 0.18\,m/s and (\textcolor[gray]{0.8}{$\blacksquare$}) 0.24\,m/s. Data were measured using the flow cell. ($\bigstar$) The wavelength of the patterns which develop when sediments are added to the system. Data were measured using the flow trough. The line shows a fit to the data for $\lambda=3.3\pm0.3h_0$, where $h_0$ is the thickness of the initial PVA film. The flow rate and viscosity have no effect on the resulting dominant wavelength that develops. (b) $\lambda/h_0$ as a function of flow speed for PVA $\eta=124$\,Pa\,s. (c)$\lambda/h_0$ as a function of viscosity for a flow speed of 0.12\,m/s.}

\caption{\label{Fig:Exp2} Time evolution of the peak to trough amplitude of ripples induced in PVA films subject to flow. (a) Amplitude as a function of film thickness for ($\medbullet$) 1mm, ($\blacksquare$) 2mm, ($\blacktriangledown$) 3mm and ($\blacktriangle$) 4mm thick films. Grey and black symbols show different runs using the same experimental conditions. Lines show the exponential fits to the data with relaxation times of between 145 and 400\,s. This corresponds to the measured relaxation time of the PVA of 115-338\,s. The variation here is seen for different batches of the PVA mixture. (b) The saturation amplitude is not dependent on the flow rate of the water. The graph shows data for 1\,mm thick films at flow rates of ($\medbullet$) 0.12\,m/s, ($\bigcirc$) 0.18\,m/s and ($\ominus$) 0.24\,m\,s$^{-1}$. The saturation amplitude is only dependent on the film thickness (c).}

\caption{\label{Fig:ExpKin} Development of lab-Kinneyia in a 2\,mm thick PVA film at a flow speed of 0.024\,m/s. Prior to flow the film is initially flat (a). The black colour comes from the underlying substrate. After 60\,s of flow sediment is deposited. Images (b) and (c) show the pattern formation 120\,s and 180\,s after the flow started. All images are 6\,$\times$\,6\,cm. The insets (2\,$\times$\,2\,cm) show the patterns at the edge of the film where it was thinner (1\,mm thick). A decrease in the wavelength is observed in the thinner film. (d) The 2D autocorrelation function of (c). The first non-zero peak is observed at 0$^{\circ}$ at a wavelength of 5.2\,mm, where$0^{\circ}$ is defined to be parallel to the flow direction. The inset in (d) shows the 2D autocorrelation function of the inset in (c). The first non-zero peak is observed at 0$^{\circ}$ for a wavelength of 2.9\,mm. (e-h) Development of lab-Kinneyia in 6\,mm thick films for flow speeds of (e) 0.024\,m/s and (f) 0.18\,m/s. (g, h) 2D autocorrelation function of (e, f). The wavelengths are 19.4\,mm and 21\,mm in (e) and (f) respectively. An increase in the order is observed with increasing flow speed. (e, f) are the same size as (a-c). The colour maps for the 2D autocorrelation functions show the intensity in arbitrary units.}

\caption{\label{Fig:FT} Photos of Kinneyia stuctures found at (a, c) Neuras farm and (b) Haruchas farm. The photographs in (a-c) have the same magnification and are 10\,x\,10\,cm. (d-f) Three-dimensional map scans taken from casts of (a-c) respectively. The scans are (d) 22.5$\times$22.5\,mm and (e, f) 55$\times$55\,mm. The inset in (d) shows the map scan in (d) at the same magnification as (e) and (f). (g-i) Two-dimensional autocorrelation functions of (d-f). The colour maps show the relative intensity of the peaks in arbitrary units. The arrows in (g) indicate the orientation of the ripples.}

32 — 1302.7280

\caption{Relative clustering error for 100 simulations with $M=2$ (a) and $M=3$ (b) data sources, shown for separate clustering (\textcolor{blue}{blue}), joint clustering (\textcolor{red}{red}), dependent clustering (\textcolor{green}{green}) and BCC (\textbf{black}). A LOESS curve displays clustering error as a function of $\alpha$ for each method. }

\caption{PCA plots for each data source. Sample points are colored by overall cluster; cluster 1 is \textbf{black}, cluster 2 is \textcolor{red}{red}, and cluster 3 is \textcolor{blue}{blue}. Symbols indicate source-specific cluster; cluster 1 is `$\bullet$', cluster 2 is `$+$', and cluster 3 is `$\ast$' . }

33 — 1303.0973

\caption{\label{fig:red.pink.white.noise}\textcolor{referee}{Summary of noise statistics for each night computed on the lightcurve residuals (data subtracted from the model) for each complete run (or equivalently for each observing night). Labels `\textsf{P}', `\textsf{R}' and `\textsf{W}' stand for \textit{pink}, \textit{red} and \textit{white}, respectively. The run (or night) number as well as the number of considered data points (\textsl{np}) is indicated above each frame. The abscissa correspond to HJD-2455000.}}

\caption{Phase-folded lightcurve of WAPS-19b, where the gray and blue dots correspond to the original data, and to the binned data at 0.001 phase units (blue), respectively. \textcolor{referee}{The black line is the best fit to the transit (see Table~\ref{Table:WASP19_TransitParam} for detailed primary transit parameters). The} \textcolor{referee2}{inset} \textcolor{referee}{frame is the normalized flux scale zoom on the secondary eclipse (full phase coverage in abscissa) with a bin size of 0.01 in phase -- see Sect.\,\ref{sect:secondary_eclipse} for detailed secondary eclipse parameters.}}

\caption{\label{fig:variations}Fluctuations of \textcolor{referee}{the phase timing (left panel), length duration (middle panel) and depth (right panel)} of individual transits of WASP-19b relative to the values from the combined transits. The crosses correspond to runs (nights) with no transit event, and orange point to nights with incomplete primary transit.}

\caption{\label{fig:snr.vs.nights}\textcolor{referee}{Standard deviation of the folded lightcurve for 1 hour bins against the number of folded nights (from best to worst). The numbers along the curve correspond to each of the newly added night number}\textcolor{referee2}{: the first point corresponds to a single night whilst the latest corresponds to the 23 nights. The lower panel shows the $\sigma_{e_k}$ value for each night.}}

\caption{\label{fig:brightness.ratio}Ratio of the brightness temperature to the maximum irradiation temperature (at the substellar point -- see text) versus wavelengths for exoplanets with detected visible and infrared secondary eclipses. Green and red curves correspond to the total modeled spectrum and to the thermal emission only, respectively. The ellipses account both for the bandpasses of the observations (horizontally) and the uncertainties of the measurements (vertically). The colors indicate whether the data originate from space telescopes (blue), mid-latitude ground based facilities (orange), or ASTEP\,400 (red).}

34 — 1303.1846

\caption{Main events of \gray emission detected by the \grid in the period November 2007 - July 2009. All detections have a significance above $3\sigma$ ($\sqrt{TS} \geqslant 3$). \textit{Column 1}: period of detection in MJD; \textit{Column 2}: significance of detection; \textit{Column 3}: photon flux (above 100 MeV).}

\caption{The $\nu F_{\nu}$ spectra of \cyg during the \gray activity. \textit{Red circles}: \grid energy spectrum (50 MeV to 3 GeV) of the main episodes. \textit{Blue error contours} and \textit{dashed blue line}: average power-law fit with $\alpha=2.70 \pm 0.25$ of the spectrum obtained by \lat integrating the two active windows of about two months each \citep{abdo_09}.}

\caption{Multiwavelength SED of \cyg during the main \gray events (non-simultaneous data) and the leptonic model \textit{``A''} (see main text). \textit{Blue circles}: X-ray average ``hypersoft'' spectrum \citep{koljonen_10}, \textit{RXTE}-PCA and \textit{RXTE}-HEXTE data ($\sim$3 to $\sim$150 keV); \textit{red circles}: \grid energy spectrum (50 MeV to 3 GeV) of the main \gray episodes; \textit{magenta arrows}: MAGIC differential flux upper limits (95\% C.L.), 199--3155 GeV, related to soft spectral state \citep{aleksic_10}. Spectral components of the model are the BB emission from the disk (blue short-dashed line), IC scattering of the soft photons from the accretion disk (green solid line), and IC scattering of the soft stellar photons (red solid line). The global SED model curve is indicated by a black solid line.}

\caption{Multiwavelength SED of \cyg during the main \gray events (non-simultaneous data) and the leptonic model \textit{``B''} (see main text). Spectral components of the model are the BB emission from the disk (blue short-dashed line), IC scattering of the soft photons from the accretion disk (green solid line), and IC scattering of the soft stellar photons (red solid line). The global SED model curve is indicated by a black solid line. For a detailed description of the datasets, see caption to Figure~\ref{cyg_x3_mw_spectrum_lep_a}.}

\caption{Multiwavelength SED of \cyg during \gray events (non-simultaneous data) and a hadronic model (see main text). Black body emission from the disk (blue short-dashed line), and \gray emission from $\pi^0$-decays (black solid line) are plotted. For a detailed description of the datasets, see caption to Figure~\ref{cyg_x3_mw_spectrum_lep_a}.}

35 — 1303.2585

\caption{\label{fig:Bint}(Colour online) Temperature dependence of the internal magnetic field $B_\mu$ (\opensquare) and of the transverse relaxation Gaussian width, $\sigma/\gamma_{\mu}$ (\fullcircle), for the investigated \fetese\single crystals.}

\caption{\label{fig:lambda1}(Colour online) Temperature dependence of the muon-spin longitudinal relaxation rate $\lambda$ for the investigated \fetese\single crystals. The light blue areas indicate the magnetic phase.}

36 — 1303.3160

\caption{ Electrochemical potential diagram for a system with 3~sites and Coulomb interaction strength~\({u_c = 2.0}\). Single particle states blue dots~(\bluesolidcircle). Two particle states red squares~(\redsolidsquare). Three particle state brown diamond~(\brownsoliddiamond).\label{fig:mu-diag}}

\caption{ Comparison of the current in the left lead with and without Coulomb interaction. Red solid line~(\redsolidsquare) has the Coulomb strength~\(u_c = 2.0\) while the blue dashed line~(\bluedashedcircle) has~\(u_c = 0.0\). The temperature is~\(k_BT_L = 0.25\) in the left lead and~\(k_BT_R = 0.10\) in the right lead. The lead-dot coupling parameters are $V_{\ell}=V_L=V_R=0.75$ and the hopping energy in the leads is $\tau=4$ (units of $t_s$). \label{fig:coul}}

\caption{ Charge as a function of the chemical potential. Brown solid line~(\brownsolidsquare) shows charging for the single particle states, black dotted lines~(\blackdottedtriangle) for two-particle states, violet dashed lines~(\violetddashedcircle) for the three-particle state, and blue dashed lines~(\bluedashedcircle) show the total charge. \textbf{(a)}~Without Coulomb interaction, \({k_BT_L = 0.25}\) and~\({k_BT_L = 0.10}\). \textbf{(b)}~With Coulomb interaction (\(u_c = 2.0\)), \({k_BT_L = 0.25}\) and~\({k_BT_L = 0.10}\). \textbf{(c)}~With Coulomb interaction (\(u_c = 2.0\)), \({k_BT_L = 0.55}\) and~\({k_BT_L = 0.40}\). \label{fig:chg-coul}}

\caption{ A comparison of the current in the left lead for two different temperatures. The red dashed line~(\reddashedcircle) has~\({k_BT_{L,R} = (0.25, 0.10)}\). While the blue solid line~(\bluesolidsquare) has~\({k_BT_{L,R} = (0.55, 0.40)}\). The temperature difference between the leads is the same in both cases~\({\Delta k_BT = 0.15}\). \label{fig:cur-temp}}

\caption{A comparison of the current created by a temperature bias with the current created by a chemical potential bias. The red solid line~(\redsolidcircle) shows the thermal current with \({k_BT_L = 0.25}\), \({k_BT_R = 0.10}\) vs. the chemical potential $\mu=\mu_L=\mu_R$. The blue dashed line~(\bluedashedsquare) shows the electrical current with no temperature bias, \({k_BT_L = k_BT_R = 0.10}\), but \({\Delta \mu = \mu_L-\mu_R=0.15}\), i.~e. energetically identical to the the previous thermal bias. On the horizontal axis we use the mean value of the left and right chemical potentials \({\mu = (\mu_L + \mu_R)/2}\). The lead-dot coupling parameters are $V_{\ell}=V_L=V_R=0.75$ and the hopping energy in the leads is $\tau=4$ (units of $t_s$).}

\caption{The blue solid line~(\bluesolidsquare) shows the current in the left lead calculated using the GME with \(u_c = 2.0\). While the red dashed line~(\reddashedcircle) shows the current calculated with the GME using eigenvectors corresponding to non-interacting electrons, but with an energy spectrum forced to be that of the system with \(u_c = 2.0\).}

\caption{A comparison of the current in the leads between systems with two and three leads. The blue dashed line~(\bluedashedcircle) shows the current for a system with two leads, as before, with $k_B T_L=0.25$ and $k_B T_R=0.10$, and the same chemical potential $\mu$. The red solid line~(\redsolidsquare) shows the current in the same system, but with a third lead attached to the middle point in the sample with $K_B T_L=0.10$ and the same chemical potential as the other leads.}

37 — 1303.3209

\caption{\textbf{Simulation error depending on initial longitudinal momentum spread.} Evolution of the mean square errors (blue $\times$) between simulations and experimental data for increasing initial longitudinal momentum spreads in the simulations, calculated from elliptical segment integration. The black solid line shows the quadratic curve fitted through the errors, and the red \fullcircle its minimum point. The red \opencircle indicate the confidence interval for the minimum position. }

38 — 1303.3751

\caption{Information Gain Values of Different Features for Different Datasets\label{table:infogain}}{ % title of Table \centering \includegraphics[width=\textwidth]{infogain_cropped.pdf} }

39 — 1303.3970

\caption{(Color online) Phonon dispersion relations for stoichiometric FeO (solid lines) and the relative intensities of the phonon modes in Fe$_{1-x}$O with $x=3$\%, for $U=6$~eV, $J=1$~eV. The experimental data (solid squares) are adopted from Ref. \cite{Kugel}. The highest intensity mode is taken as a reference (100\%); relative intensities are arranged into the following ranges indicated by the intensity of gray scale (color): 5-20\% (\textcolor{lightblue}), 20-50\% (\textcolor{cyan}), 50-80\% (\textcolor{blue}), 80-100\% (\textcolor{red}). High-symmetry points are (in units of $2\pi/a$): $\Gamma=(0,0,0)$, $X=(\frac12,\frac12,0)$, $K=(\frac38,\frac38,\frac34)$, $L=(\frac12,\frac12,\frac12)$, $W=(\frac12,\frac14,\frac34)$. }

40 — 1303.5057

\caption{Results of the grid projection benchmark for the configuration shown in the lower-right panel of Figure~\ref{fig:projections}. The run times of three different algorithms, \Geokerr\(blue diamonds), \Ray\(green triangles), and \GRay\(red circles) in double precision, are plotted against the number of geodesics traced for each image. The asymptotic linear dependence seen for all three algorithms demonstrates explicitly that the ray tracing problem is highly parallelizable. For a small number of geodesics, the performance of \GRay\flattens to a constant value (approximately 20~ms for the configuration used) because of the time required for launching the CUDA kernel.}

41 — 1303.5199

\caption{Approximated $\epsilon_{app}$ ('\textcolor[rgb]{0.00,0.00,1.00}{-}') and 400 randomly generated samples of $\theta$. '\textcolor[rgb]{1.00,0.00,0.00}{$\ast$}' represents samples which are satisfying $V_{app}(\theta) \leq \frac{1}{\gamma}$ while '$o$' are those located outside the application set. 81\% of the samples that fulfill the condition $V_{app}(\theta) \leq \frac{1}{\gamma}$ are located inside the approximated ellipsoid given by the proposed method.}

\caption{Approximated ('\textcolor[rgb]{0.00,0.00,1.00}{o --}') and real ('\textcolor[rgb]{1.00,0.00,0.00}{. - -}') values of $V_{app}(\theta)$ for 100 different samples of $\theta$ taken form a uniform distribution.}

\caption{Approximated ('\textcolor[rgb]{0.00,0.00,1.00}{o --}') and real ('\textcolor[rgb]{1.00,0.00,0.00}{. - -}') values of $V_{app}(\theta)$ inside the application set. 92\% of the samples inside the region are classified as acceptable ones by the proposed method.}

42 — 1303.5790

\caption{Left: Dependence of the gain with the amplification field. The empty marker represents the maximum gain reached for each pressure, in case it was not determined for the given conditions but for a slightly different gas mixture around the optimum. Right: Dependence of the best energy resolution with pressure at 22.1 keV, for data from \cite{Coimbra} (\textcolor{green}{$\blacktriangledown$}), from \cite{XeZaragoza} (\textcolor{red}{$\blacktriangle$}) and from this work (\textcolor{blue}{$\blacksquare$}). % due to a detector's limitation by dust. %Limitations on the maximum gain were overcome by washing the detector with water and then flushing with helium. \label{fig:GGainvsAmFi}}

43 — 1304.1282

\caption{\label{fig:Vphasespace} Contour plots in the tetragonal parameter space: (a) total energy landscape and (b) $N(E_{\rm F})$ of vanadium. The two minima of $E(a,c/a)$ are indicated by ${\color{blue} \blacksquare}$ and the saddle point at the FCC structure is indicated by {\color{blue} $\times$}. The inset shows the saddle point region in more detail. The global minimum corresponds to the stable BCC phase, the local minimum is a BCT structure. The EBP (black dot-dashed line) has two discontinuities, indicated by black dashed lines. Brown hatched lines signal axial ratios of the cubic structures. The energy and the density of states were calculated on a uniform $(a,c)$ grid with a distance of $0.025\,\textrm{\AA}$ between two adjacent points.}

44 — 1304.3804

\caption{Function {\tt buf\_flush\_buffered\_writes} of {\tt MySQL}: worst-case running time plots with curve fitting.}

45 — 1304.4299

\caption{(Color online) (a) Level diagram showing all 52 possible states addressed by the experiment. The arrows indicate allowed excitations for $\sigma$ polarized probe and coupling beams and $\pi$ polarized MW's. The $54P_{3/2}$ states are shown above the $53D_{5/2}$ states for simplicity. On the right, the corresponding effective 4-Level system is shown. (b) Theoretical lineshapes resulting from a 3-level (black) and 4-level (\textcolor{red}{red}) system.}

\caption{(Color online) Theoretical (a) and experimental (b) results for the illustrative polarization cases described in the text: probe laser, coupling laser and MWs $\hat{x}$-polarized (black); probe and coupling laser $\hat{y}$-polarized and MWs $\hat{x}$-polarized (\textcolor{blue}{blue}); probe and coupling lasers $\sigma^+$ polarized and MWs $\hat{z}$-polarized (\textcolor{red}{red}). The additional line broadening observed in the experiment is due to the MWs being inhomogeneous over the extent of the vapor cell. The effect resulted from the positioning of the antenna that was required to avoid unwanted reflections in our laboratory.}

46 — 1304.5193

\caption{\label{fig:model} Two-dimensional representation of the flux-tube model. The solid lines mark the magnetic field lines on a $x$-$z$ plane at {$y=0$~Mm}. The $\beta=1$ (\dashed) and $\beta=0.1$ (\dotted) contours are also shown.}

47 — 1304.5304

\caption{ Example network realization. The reference receiver (indicated by the {\color{red} $\blacktriangledown$}) is placed at the origin, and the reference transmitter is at $X_{0}=1/6$ (indicated by the {\color{red} $% \bullet$} to its right). $M=30$ mobiles are placed according to the uniform clustering model, each with an exclusion zone (not shown) of radius $r_% \mathsf{ex}=1/12$. Active mobiles are indicated by filled circles, and deactivated mobiles are indicated by unfilled circles. A guard zone of radius $r_\mathsf{g}=1/4$ surrounds each active mobile, as depicted by dashed circles. When CSMA guard zones are used, the other mobiles within the guard zone of an active mobile are deactivated. \protect\vspace{-0.65cm} }

48 — 1304.5513

\caption{Electrical conductivities calculated with GeoDict. The solid squares ({\color{red}{$\blacksquare$}}) are the values estimated by using \citet{Arora2004}, and the solid line is from the Archie's law \cite{Achie} $\sigma_{zz}/\sigma_2=(1-q)^\alpha$ using $\alpha=2.35$.}

\caption{Simulation results from in-house developed routines for Young's modulus as a function of fiber concentration. The solid lines correspond to $q^2$ and $q^6$ relationships. The solid squares ({\color{red}{$\blacksquare$}}) are values from GeoDict and the solid diamonds are from experimental data.)}

\caption{Effective conductivity of a fiber and electrolyte system as a function of fiber concentration. Simulation results from in-house routines are shown with open symbols. Results from GeoDict are shown with solid squares ({\color{red}{$\blacksquare$}}). The solid lines that are used to define the bounds are estimated for particle filled composites using the Bruggeman expression \cite{Bruggeman1935} with shape parameters 1.8 and 4 for lower and upper bounds, respectively. The other lines that are drawn to represent the data are for Archie's law with $\alpha=2.35$ (estimated for GeoDict data) and $\alpha=2$ which represents the in-house results precisely. The inset is the same plot in logarithmic concentration axis. }

49 — 1304.5873

\caption{ Numerical simulations of Eq.\\ref{kpz} ($\bullet$) and Eq.\\ref{fake} for $\mu=3/2$ ({\color{OliveGreen}$\scriptstyle{\blacksquare}$}) and $\mu=7/4$ ({\color{red}$\blacklozenge$}). Surface structure factor (upper row) and roughness (lower row) for $d=1$ (left column) and $d=2$ (right column). Solid and dashed lines represent power-law behaviors as indicated, which are discussed in the main text. All the observables have been averaged over $10^3$ ($10^2$) different noise realizations for $d=1$ ($d=2$), starting from a flat initial condition. Error bars are smaller than symbol sizes. All units are arbitrary.}

\caption{1D Height distributions for Eq.\\ref{kpz} ($\bullet$) and Eq.\\ref{fake} for $\mu=3/2$ ({\color{OliveGreen}$\scriptstyle{\blacksquare}$}) and $\mu=7/4$ ({\color{red}$\blacklozenge$}). The variable $\chi$ is defined in the text. The solid blue line is the TW (GOE) distribution expected for $d=1$ \cite{edelman:2008}. $P(\chi)$ is estimated from 2048 independent runs starting from a flat initial condition. Inset: zoom of main panel, in linear representation. All units are arbitrary.}

\caption{ 1D height-height correlation function for Eq.\\ref{kpz} ($\bullet$) and Eq.\\ref{fake} for $\mu=3/2$ ({\color{OliveGreen}$\scriptstyle{\blacksquare}$}) and $\mu=7/4$ ({\color{red}$\blacklozenge$}). Here, $u= x\sqrt{\Gamma/2\lambda}/(2\Gamma t^*)^{2/3}$ while $C(x,t=t^*)$ is measured from the same surfaces employed to estimate $P(\chi)$ in Fig.\\ref{fig:2}. The solid blue line provides the covariance of the Airy$_1$ process \cite{bornemann:2010}. All units are arbitrary.}

50 — 1304.6850

\caption{\label{predictions} Model predictions for $\rho^0$ photoproduction at central rapidities ($y=0$), for the models GGM \cite{ggm1, ggm2}, shown with a full (\full) red line, GM \cite{gm}, shown with a yellow chained (\chain) line and Starlight \cite{starlight, starlighturl}, shown with a blue dashed (\dashed) line. (Color online.)}

\caption{\label{fig:bwfit}The invariant mass distribution fitted with a Breit--Wigner function with continuum correction. Linear (left) and logartimic (right) scale. The data points are marked with a full circle (\fullcircle), the blue dashed line (\dashed) is the resonance plus the continuum, the full green line (\full) is the resonance plus the continuum in the range used to make the fit and the red dotted line (\dotted) is only the resonance. Statistical errors are shown. (Color online.)}

\caption{\label{fig:pt} The transverse momentum distribution compared to Starlight predictions for coherent and incoherent photoproduced $\rho^0$'s. The data points are marked with a full circle (\fullcircle), the simulated coherent production is marked with a red full line (\full), the simulated incoherent production is marked with a full blue line (\full), and the sum of the incoherent and coherent simulation is marked with a dashed green line (\dashed). (Color online.)}

\caption{\label{fig:zdcenergy}Energy deposited in the ZDC on the A--side (left) and the C--side (right). Each peak is fitted with a Gaussian. The first peak represent zero neutrons, the second peak one neutron and so on. Because a pedestal value is subtracted from the signal, the energy goes below zero. The data points are marked with a full circle (\fullcircle), the fit to each peak each marked with a green dashed line (\dashed), and the sum of the fits is marked with a red full line (\full). (Color online.)}

51 — 1304.6991

\caption[Centerline velocities plotted for lid-driven cavity]{(Top) Streamfunction and Pressure contours with a single spectral element of order 16. (Bottom) Centerline velocities are plotted ({\color{black}{\solidrule}}) and compared with the solutions of \citep{Botella} ({\color{red}{\solidrule}}), and the solutions are found to be reasonably close. Mesh size is $4\times 4$ and made up of elements of order 6.}

52 — 1304.7917

\caption{\footnotesize Shapes of the potential for some values of $E$\,. The critical electric-field is denoted by$E_{\text{c}}$\,.\textcolor{blue}{The blue line} is the barrier with $E<E_{\text{c}}$\,.\textcolor{red}{The red line} with $E=E_{\text{c}}$\,.\textcolor{green}{The green line} with $E>E_{\text{c}}$\,.}

53 — 1305.0376

\caption{Current $j_L(r_1,r_2,\alpha,\beta)$ obtained by Monte Carlo simulations on a lattice of $L=1000$ sites with $15$ slow sites of two types, $10$ slow sites with rates $r_1$ (type 1) and $5$ slow sites with rates $r_2$ (type 2), plotted as a function of $r_1$ for fixed $r_2=0.3$ (\full) and $r_2=1$ (\opencircle). Setting $r_2=1$ in the latter case means that only type 1 slow sites are present. Broken line (\broken) corresponds to $r_1=1$ and $r_2=0.3$ (type 2 slow sites only). In all cases $\alpha=\beta=1$. Slow sites of type $1$ were placed at $i=27,79,262,558,563,655,686,701,720,888$ and those of type $2$ at $i=156,191,407,744,997$.}

54 — 1305.0703

\caption{\leg{Wall pressure} vs. packing fraction: (\textcolor{blue}{$\bigcirc$}) : $P_{TOT}$, (\textcolor{red}{$\square$}): $P_{STAT}$, (\textcolor{green}{$\triangle$}): $P_{DYN}$, as defined in the text for the present PSM-4 disks experiment. Note: the finite stiffness of the piston has been calibrated and removed from the data. The vibration frequency is $f=10$~Hz.}

\caption{{\bf relaxation dynamics of the contact network.} (color online). \leg{(a):} Temporal average of the overlap function $Q^z(\tau)$. Packing fractions as in figure~\ref{fig:glass}. \leg{(b):} Plateau value of $Q_s^z$, defined by $Q_s^z=Q^z(\tau=4)$ (\textcolor{red}{$\blacktriangle$}), and fraction of non-rattling particles (\textcolor{blue}{\bf\texttimes}), vs. packing fraction $\phi$. \leg{(c):} Relaxation time of the contact network, $\tau_\alpha^z$ (see text for definition) (\textcolor{red}{$\blacksquare$}, left axis), and average contact number, $z$ (\textcolor{blue}{\bf\texttimes}, right axis), vs. packing fraction, $\phi$. The plain red line is a fit of the form $\tau_\alpha^z\sim (\phi^\dagger-\phi)^{-2.0}$. The dashed line indicates $\phi^\dagger=0.8151$. The vibration frequency $f=10$~Hz.}

\caption{{\bf Motion of the center of mass.} (color online) \leg{(a)} Center of mass position, in the vibration direction $X_b$ \leg{(top)}, and in the transverse direction $Y_b$ \leg{(bottom)}, vs. time $t$, at $\phi=0.8089$ (blue), $0.8161$ (green) and $0.8196$ (red). \leg{(b):} Amplitudes $A_b^X$ (\textcolor{blue}{$+$}) and $A_b^Y$ (\textcolor{red}{$\bigcirc$}) vs. packing fraction, $\phi$. The vibration frequency $f=10$~Hz. }

\caption{{\bf Mean square displacements.} (color online). \leg{(a):} Mean square displacements $MSD$ for filtered trajectories (see text) vs. lag time $\tau$ for both the short time (fast camera) and long time (stroboscopic acquisition) experiments. The packing fractions explore the same range and are color coded as in figure~\ref{fig:glass}, the binning being finer for the fast camera acquisition. \leg{(b):} Plateau value $\Delta^2$ obtained from the short time data MSD (\textcolor{green}{$+$}), from the long time stroboscopic data MSD (\textcolor{blue}{$\times$}), and from the low frequency limit of Energy Spectral Density, ${\Delta^2}^0_f$ (\textcolor{red}{$\square$}) vs. reduced packing fraction, $\epsilon=(\phi-\phi^\dagger)/\phi^\dagger$. \leg{(c):} Plateau entrance time $\tau_{En}$ vs. reduced packing fraction $\epsilon=(\phi-\phi^\dagger)/\phi^\dagger$. The vibration frequency $f=10$~Hz, i.e $\gamma=1.4$.}

\caption{{\bf Dynamical correlations of the contacts at short and long times.} (color online) \leg{(a):} Dynamic susceptibility of the contacts $\chi_4^z$ vs. the lag time $\tau$. Same packing fractions as in figure~\ref{fig:glass}. \leg{(b):} Dynamic susceptibility at short time ${\chi_4^z}^0$ (\textcolor{blue}{\bf\texttimes}) and maximal dynamical susceptibility ${\chi_4^z}^*$ (\textcolor{red}{\bf$\Box$}) vs. reduced packing fraction, $\epsilon$. The vibration frequency $f=10$~Hz, i.e. $\gamma=1.4$.}

\caption{{\bf Hard vs. Soft.} (color online) Piston force (top) and Maximal dynamical susceptibility of the displacements (bottom) vs. reduced packing fraction, $\epsilon$, for \leg{(a)}: hard brass disks ~\cite{lechenault_epl1} and \leg{(b)}: soft photo-elastic disks. (\textcolor{blue}{$\bigcirc$}). $P_{TOT}$, (\textcolor{red}{$\square$}): $P_{STAT}$, (\textcolor{green}{$\triangle$}): $P_{DYN}$ as in figure~\ref{fig:P_vs_phi}. The vibration frequency $f=10$~Hz, i.e. $\gamma=1.4$. Dashed lines indicate $\epsilon^*$ and $\epsilon=0$.}

\caption{{\bf Towards zero vibration} (color online) \leg{(a):} Maximum of the dynamical susceptibility of the contact maximum$({\chi_4^{z}}^*)$ for soft grains (\textcolor{red}{$\times$}) and hard grains (\textcolor{vert}{$\bigcirc$}), estimated from max$({\chi_4^{\vec r}}^*)/20$, versus the split $|\epsilon^*|$ between static and dynamics signatures of Jamming. \leg{(b):} MSD Plateau vs. density $\epsilon$, for $\gamma=0.5$ (\textcolor{red}{$\bigcirc$}), $\gamma=0.8$ (\textcolor{vert}{$\times$}), $\gamma=1.4$ (\textcolor{bleu}{$\square$}) and for hard brass disks at $\gamma=1.4$ (\textcolor{pink}{$\Diamond$}).}

55 — 1305.1189

\caption{\label{rad} A, B, C.--Huygens' constructions: A in a \textcolor{ForestGreen}{continuous medium}; B in a \textcolor{red}{low pressure gas}; C \textcolor{blue}{including rare molecules} possibly made up by two colliding molecules.}

\caption{Addition of sine wave of different frequencies: While two sine curves \textcolor{red}{\bf *} \textcolor{blue}{*} have a constant amplitude, their sum\textcolor{violet}{*} shows the beginning of a beat.}

56 — 1305.1495

\caption{\small Comparison of test-cases used for several tracking algorithms, quantified in terms of temporal duration $T$ and estimated number of tracked objects $N$. The largest datasets processed by different tracking algorithms define the Pareto frontier in the two-dimensional space $\lbrace T,~N\rbrace$. The points are classified according to the field of investigation for which the respective algorithms have been developed: fluid dynamics experiments ({\tiny$\blacksquare$}), biological experiments ($\bullet$), and the experimental data processed using GReTA and presented in this paper and listed in Table~\ref{tab:exp} (\textcolor{red}{$\triangle$}~and~\textcolor{red}{$\bigtriangledown$} for birds and insects, respectively). The numbers next to the symbols correspond to the references to the papers from which $T$ and $N$ have been estimated -- see Bibliography.}

57 — 1305.3057

\caption[aaa]{Modal densities observed on one piano soundboard (dots, data taken from~\cite{Ege2013a}) and evaluated with the model proposed in this article (lines). The estimated values are the reciprocal of the moving average of six successive modal spacings, and reported at the mid-frequency of the whole interval.\CR Observed values at points \textbf{A}$_\mathbf{1}$ ({\color[rgb]{0,0,1}\tiny{$\bullet$}}), \textbf{A}$_\mathbf{2}$ ({\color[rgb]{1,0,0}$\blacktriangle$}), \textbf{A}$_\mathbf{3}$ ($\blacktriangledown$), and \textbf{A}$_\mathbf{5}$ ({\color[rgb]{0,0.5,0}\scriptsize{$\ast$}}), whose locations are given in \Fig{board_threefaces}~(b). The choice for averaging explains why the first estimated point is well above the first detected mode of the soundboard, at 114~Hz.\CR Sub-plate model in the low-frequency regime (\S~\ref{sec:LowFreq}): gray lines.\CR{\color[rgb]{.5,.5,.5}\hdashrule[0.5ex]{5em}{1pt}{2.5mm 1mm .3mm 1mm}}~: "Norway spruce". {\color[rgb]{.5,.5,.5}\hdashrule[0.5ex]{5em}{1pt}{2.5mm 1mm}}~: "Sitka spruce".\CR{\color[rgb]{.5,.5,.5}\hdashrule[0.5ex]{5em}{1.5pt}{.3mm .5mm}}~: "Mediocre spruce" (see Table~\ref{tab:caracmeca_num} for parameter values). Wave-guide(s) model in the high-frequency regime (\S~\ref{sec:HighFreq}), with Norway spruce parameters: colored lines.\CR Lower {\color[rgb]{0,0,1}\hdashrule[.5ex]{4em}{1pt}{}} line~: modal density of the (1,$n$)-modes in one single inter-rib space enclosing point \textbf{A}$_\mathbf{1}$ (see Eq.~(\ref{eq:ModalDenWG})); the horizontal line at the right hand-side of the figure corresponds to the asymptotic value of the modal density in this wave-guide, with all possible $(m,n)$-modes.\CR Group of thin lines: modal density in the set of three adjacent wave-guides. {\color[rgb]{0,0,1}\hdashrule[0.5ex]{4em}{1pt}{}}~: set of wave-guides in the vicinity of \textbf{A}$_\mathbf{1}$; {\color[rgb]{1,0,0}\hdashrule[0.5ex]{4em}{1pt}{2.5mm 1mm}}~: vicinity of \textbf{A}$_\mathbf{2}$.\CR{\hdashrule[0.5ex]{4em}{1pt}{0.3mm 0.5mm}}~: vicinity of \textbf{A}$_\mathbf{3}$;\quad{\color[rgb]{0,.5,0}\hdashrule[.5ex]{5em}{1pt}{2.5mm 1mm .3mm 1mm}}: vicinity of \textbf{A}$_\mathbf{5}$.\CR Thick {\color[rgb]{0,0.5,0}\hdashrule[.5ex]{4em}{1pt}{}}~line: transition between the three-waveguide model and the sub-plate model (\S~\ref{sec:Transition}), for point \textbf{A}$_\mathbf{5}$. }

58 — 1305.3389

\caption{Dispersion relation and determination of the spontaneous symmetry breaking process. The vertical solid lines represent the values of $K_{11}/B$ at $h=0.7$ and $R=1$, for two different values of $B$: (\textcolor{red}{---}) $B=5$, (\textcolor{green}{---}) $B=1$. The thick blue line displays $BD(K)$, given by Equation~\ref{DBK}. Their intersections yield the $D_{11}$ where zero modes appear and the first spontaneous symmetry breaking solution occurs.}

\caption{Test of the model for the turbulent von K\'arm\'an experiment: (\textcolor{blue}{---}) $K$ vs ${\rm Re}$ after the mapping of Section~\ref{mapping}. The red line indicates $K_{11}$. Its intersection with $K({\rm Re})$ occurs at ${\rm Re}\approx40\,000$. The three couples of insets show the comparison between theoretical velocity fields (right), computed using the out-of-equilibrium toy model, and experimental velocity field(left) at ${\rm Re}=5000$, ${\rm Re}=43\,000$ and ${\rm Re}=200\,000$. The color codes the azimuthal field, the arrows indicate the $u_r,u_z$ field. The red dashed lines correspond to the $z = 0$ plane.}

59 — 1305.3556

\caption{Effect of control parameters on the thermodynamic optimal length and charge ratio. {\bf (A)} Effect of increasing capsid charge, with capsid $\inrad=7.3$ nm. {\bf (B)} Effect of increasing capsid size for fixed ARM length=5. {\bf (C)} Effect of base-pairing, with $\fbp=0.5$ base-paired nucleotides and varying maximum ladder distance (MLD), for capsid $\inrad=7.3$ nm and ARM Length=5. Snapshots of our model NA structures with small and large MLD's are shown (prior to encapsidation), with double-stranded regions in red and single-stranded regions in blue. The result for no base-pairing (linear) is shown as a dashed line. {\bf (D)} Effect of ionic strength and comparison between Debye-Huckel interactions and explicit ions. The thermodynamic optimum lengths $\Lopt$ and corresponding optimal charge ratios are shown as functions of the ionic strength (Debye screening length), calculated with simulations using Debye-Huckel (DH) interactions (\textcolor{red}{{\Large$\bullet$}} symbols) or Coulomb interactions with explicit ions (\textcolor{blue}{\DiamondSolid} symbols). \label{fig3} }

\caption{Comparison between viral genome lengths and calculated thermodynamic optimal lengths {\bf (A)} and charge ratios {\bf (B)} for models based on the indicated viral capsid structures. Predicted optimal lengths are shown for linear polyelectrolytes ({\Large\textcolor{red}{$\bullet$}} symbols) and model NAs (\textcolor{blue}{$\blacktriangle$} symbols) with 50\% base-pairing. Viral genome lengths are shown with \textcolor{OliveGreen}{\pentagon} symbols. Error bars fall within the symbol sizes. \label{fig4} }

\caption{Number of NA segments that directly interact with positively charged ARM segments (interaction energy $\leq -0.5 \kt$, \textcolor{blue}{$\blacksquare$} symbols) and bridging segments (interaction energy $> -0.5 \kt$, {\Large\textcolor{Orchid}{$\circ$}}). The numbers are calculated at the optimal length $\Lopt$ for each capsid shown in Fig.~\ref{fig4} using the base-paired model. For visual reference, the dashed line indicates a 1:1 correspondence between capsid charge and number of nucleotides. \label{fig5} }

\caption{{\bf (A)}Effect of ionic strength and comparison between Debye-Huckel interactions and explicit ions. The thermodynamic optimum lengths $\Lopt$ and corresponding optimal charge ratios are shown as functions of the Debye screening length (bottom axis) and ionic strength (top axis), calculated with simulations using Debye-Huckel (DH) interactions ({\Large\textcolor{red}{$\bullet$}} symbols) or Coulomb interactions with explicit ions, 1:1 salt and no divalent cations (\textcolor{blue}{\DiamondSolid} symbols), 5\% 2:1 salt (\textcolor{purple}{\TriangleDown} symbols) or 10\% 2:1 salt (\textcolor{green}{\TriangleUp} symbols). An additional system with monovalent free ions and divalent cations irreversibly bound to the polyelectrolyte is also presented ({\Large\textcolor{orange}{$\diamond$}} symbols). Calculations were performed using the simple capsid model (Fig.~\ref{fig1} in the main text) and a linear polyelectrolyte. All ion radii were set to 0.125 nm. {\bf (B)}Effect of varying the ion radius ({\Large\textcolor{blue}{$\bullet$}} symbols) at 100 mM monovalent ions, compared with the DH result (dashed red line). \label{figS1} }

\caption{{\bf (A)} Schematic illustrations of the algorithm we used to obtain a wide range of base-paired structures. From left to right, double-stranded (ds) segments are first randomly assigned. These segments are then base-paired together, starting from one end. If base-paired segments are widely separated (i.e. $\lpair$ is large) then subsequent nested base-pairs lead to an extended structure. Conversely, if $\lpair$ is small, less extended structures form. The right-most panel indicates a psuedoknot, a structural motif we have prevented from occurring in this model, by setting $l_\text{max}$ to the last unpaired segment. {\bf (B)} Radius of gyration $\Rg$ for model NAs isolated in solution as a function of maximum ladder distance (MLD) normalized by the maximum possible MLD. The nucleic acid has 1000 nt, 50\% of which are base-paired. {\bf (C)} The frequency of junction numbers can be altered by varying $\lambda$ in Eq.~\ref{eq:lPair}, with large values of $\lambda$ leading to large values of $\lpair$. The symbols indicate the relative frequency of junction numbers for biological RNAs with indicated lengths, obtained from Ref.~\cite{Gopal2012}, and the lines are best fits to these distributions generated by varying $\lambda$. The inset illustrates several different junction orders. {\bf (D)} The thermodynamic optimum length measured for the simple model capsid as a function of the fraction of base-paired nucleotides $\fbp$ for a simplified `hairpins only' model (\textcolor{red}{$\blacksquare$} symbols). {\bf (E)} Snapshots illustrating assembly around a NA. Beads are colored as follows: blue=excluders, yellow=ARM bead, red=single-stranded NA, cyan=double-stranded NA. `Top', `Bottom', and `Attractor' beads removed for clarity. \label{figS3} }

\caption{Structure of the encapsidated genome. {\bf (A,D)} Radial density for linear polymer and ARM segments in the simple capsid (A) and CCMV (D). {\bf (B,E)} Angular density of linear polymer segments (heat map) in the basic capsid model (B) and CCMV (E). Green squares indicate the first ARM segment. Segment densities are averaged over radial distances of 5-6.25 nm (B) and 8.75-10 nm (E), as a function of the spherical angles. {\bf (C)} Number of polymer segments interacting with positive capsid charges (\textcolor{red}{$\blacktriangledown$} symbols), and number of polymer segments not interacting with positive charges (bridging segments, \textcolor{blue}{\DiamondSolid} symbols). The numbers are calculated at the optimal polymer length $\Lopt$ as a function of capsid inradius $\inrad$ for the simple capsid with constant ARM length. {\bf (F)} Number of NA segments interacting with positively charged ARM segments and bridging. The numbers are calculated at the optimal length $\Lopt$ for each capsid using the linear and base-paired model. The 1:1 line is plotted for reference.}

\caption{The residual chemical potential $\ur$ calculated by the Widom test-particle insertion method is shown for a linear polyelectrolyte, isolated in solution (\textcolor{red}{$\blacksquare$} symbols) and encapsidated in the simple capsid model (Fig.~\ref{fig1} main text) ($\bullet$ symbols). Results from replica exchange (REX) simulations on the encapsidated polymer are also shown (\textcolor{blue}{$\blacktriangle$} symbols).}

\caption{The optimal charge ratio as a function of polyelectrolyte linear charge density predicted by simulations in which electrostatics were calculated using Debye-Huckel (DH) electrostatics with no assumed counterion condensation (\textcolor{red}{$\bullet$} symbols) or Coulomb interactions with explicit counterions (\textcolor{blue}{\DiamondSolid} symbols). The simulations used the simple capsid model ($\inrad=7.3$, ARM Length = 5) at a Deybe length of $\ld=1$ nm or a counterion concentration of 100mM, respectively. The results are compared to the results of simulations with the Debye-Huckel model and irreversible counterion condensation (black line).}

60 — 1305.5075

\caption{The average blue (top) and red (bottom) spectra {\color{blue}of} CI Aql, uncorrected for orbital motion.}

\caption{Trailed {\color{blue}spectra} of the $\lambda\lambda5150-5290$\AA\region, showing absorption features{\color{blue} (lighter grey-scales)} of Mg {\sevensize I}b, Ca {\sevensize I} and Fe {\sevensize I} from the secondary. The data have been folded in orbital phase for clarity. Note that the noisier data visible around phases 0.1--0.4 are due to a period of cloud. This is also evident in Figs 4, 6, 9 \& 10.}

\caption{Radial-velocity curve of He {\sevensize II} $\lambda4686$\AA\measured using a double-Gaussian fit with a separation of$ 900 $ km\,s$^{-1}$. The data have been folded for clarity. The horizontal dashed line represents the systemic velocity.}

61 — 1305.5136

\caption{\label{fig:JNG}Group hierarchies of \JNG network with (b)~\logl{-500.9} and (c)~\logl{-445.3}, where node shapes correspond to high-level packages of JUNG network library, i.e., \texttt{graph} ({\color{blue} squares}), \texttt{io} ({\color{mocha} triangles}), \texttt{algorithms} ({\color{asparagus} diamonds}), \texttt{visualization}~({\color{gray} circles}) and other~({\color{black} pentagons}). (Shades of the inner nodes of the hierarchy are proportional to probabilities $\probability$.)}

62 — 1305.5171

\caption{\coloronline The LDOS in the vortex in the Born limit ($\delta_0=0$) at $T/T_{\text{c0}}=0.3$ and $\Gamma_\text{n}/\Delta_0=0.3$. left: $l_z=0$, right: $l_z=2\hbar$}

\caption{\coloronline The LDOS in the vortex in the unitary limit ($\delta_0=\cpi/2$) at $T/T_{\text{c0}}=0.3$ and $\Gamma_\text{n}/\Delta_0=0.3$. left: $l_z=0$, right: $l_z=2\hbar$}

\caption{\coloronline The dependence of the peak value of zero energy LDOS at the center of the vortex upon the scattering phase-shift $\delta_0$. The temperature $T$ and scattering rate $\Gamma_\text{n}$ are $T/T_{\text{c0}}=0.1$ and $\Gamma_\text{n}/\Delta_0=0.3$ (a), $T/T_{\text{c0}}=0.3$ and $\Gamma_\text{n}/\Delta_0=0.3$ (b), $T/T_{\text{c0}}=0.1$ and $\Gamma_\text{n}/\Delta_0=0.4$ (c) and $T/T_{\text{c0}}=0.3$ and $\Gamma_\text{n}/\Delta_0=0.4$ (d). The arrows indicate $\delta_0$ where the energy gap on the Fermi level in the bulk disappears. The critical values are numerically calculated using \eqref{EQ:gap-closing-phase-shift}.}

\caption{\coloronline The minimum excitation energy in the bulk.}

63 — 1305.5736

\caption{Measured instability threshold $Q_m$, for levitated drops. The upper axis gives the gas velocity, estimated by dividing the total flow rate by the area of the porous medium. Data represents all data points for water- and water/glycerine drops, in circles (\textcolor{blue}{$\circ$} and \textcolor{blue}{$\bullet$}) and squares (\textcolor{red}{$\square$} and \textcolor{red}{$\blacksquare$}), respectively. Since for the smallest flow rate the drop size could not be measured (it detaches from the needle), $R$ is measured in levitated state instead of sessile state. These points are therefore indicated by a solid symbol (\textcolor{blue}{$\bullet$} and \textcolor{red}{$\blacksquare$}). Note that point \textcolor{red}{$\blacksquare$} corresponds to the chimney instability from Fig.~\ref{fig:Breathing}b. The theoretical prediction of the critical radius for chimney instability is indicated by the blue dashed line and red dotted line for the used water and water-glycerine mixture, respectively.}

\caption{The frequency measured for faceted drops as shown in the images of Fig.~\ref{fig:ExampleModes}. Each data point \textcolor{blue}{$\circ$}, corresponds to one water drop measurement. The red solid line is the prediction from the corresponding eigen mode for a puddle, given by eq. (\ref{eq:cylindricalRayleigh}).}

64 — 1305.5812

\caption{Positions of the observatories for the Dst ($\bigstar$) and the Kp/ap indices (\textcolor{gray}{$\blacksquare$}).}

65 — 1305.6798

\caption{\label{fig:sisj}(Color online) The magnitude of the spin-spin correlation function $\left|\langle S_i^z S_j^z\rangle\right|$ at half-filling as a function of distance, $d/a$, for $T/t=0.2$ (circles) and 1.0 (squares), for a 16-site (black online) and 50-site (red online) cluster. For $T=1.0$ the extrapolation to the TL is shown, with exponential fit $|\langle S_i^z S_j^z\rangle| = A e^{-d/\xi}$ with $\xi\approx0.43$ and $A=\langle{S_i^z}^2\rangle\approx 0.78$. } \end{figure} % %\begin{figure} % \begin{center} % \includegraphics[width=0.95\linewidth]{lambda-cut-fits-twoframe.eps} % \end{center} % \caption{\label{fig:lambda}(Color online) The spin-spin correlation length, $\xi$ as defined in Eq.~(\ref{eqn:lambda}), in units of the lattice spacing, $a$, as a function of temperature for 20 and 50-site clusters along with extrapolation to the TL. Highlighted region marks the pseudogap onset near $T^*$. } %\end{figure} % Next we discuss the reduction of temperature towards the pseudogap onset, $T^*$. One expects that as temperature is lowered the length scale of correlations in the system should grow. To ensure that our clusters have sufficient size to account for this increasing correlation length, we increase the cluster size until we see convergence in a quantity of interest. \begin{figure} \begin{center} \includegraphics[width=0.95\linewidth]{Fig2-maydata-notl-fixsisj.eps} \end{center} \caption{\label{fig:D}(Color online) Double occupancy, $D(T)$, energy, $E(T)$, and nearest neighbour spin correlations at half-filling are shown in frames (a), (b) and (c) respectively as functions of $T/t$ at half-filling for $U/t=8$. Results are shown for 20 (black online) and 50 (red online) site cases as well as extrapolations to the TL (green online) as described in the text. Inset of (a) and (b) are enlargements of the low temperature regions of their respective figures. The inset of (b) includes extrapolated NLCE data.\cite{ehsan:priv, khatami:2011} } \end{figure} The spin-correlation function, $\langle S_{i}^z S_j^z \rangle$, as a function of distance, $d=|\boldsymbol{x}_i-\boldsymbol{x}_j|$, is such a quantity. Since the system is antiferromagnetic we remove the alternating sign and instead plot the magnitude for each neighbour distance, $\left|\langle S_{i}^z S_j^z \rangle \right|$ in Fig.~\ref{fig:sisj}. As can be seen at high temperature, $T/t=1.0$, there is little variation between a 16-site and 50-site calculation of spin correlations, shown as black and red squares. One can see that all relevant correlations are accounted for, as the amplitude of $\left| \langle S_{i}^z S_j^z \rangle \right|$ decays to zero within the linear cluster size of the small cluster. Also shown is the extrapolation of $\left|\langle S_{i}^z S_j^z \rangle \right|$ to the thermodynamic limit for $T/t=1.0$, which can be reasonably fit by an exponential decay $\left| \langle S_{i}^z S_j^z \rangle \right| = \langle {S_{i}^z}^2\rangle e^{-d/\xi}$ as expected from analytic work on the 2D Hubbard model.\cite{auerbach} The spin-spin correlation length fitting results in $\xi\approx 0.43$ which is smaller than half the linear cluster size for both the 16 and 50-site cases. Thus there is no new physics in the 50-site case which is not in the 16-site case at $T/t=1.0$. Any difference in results for thermodynamic properties for increasing cluster size must be perfectly accounted for by the $1/N$ DCA scaling. The utility of this analysis becomes apparent at low temperatures, illustrated here at $T/t=0.2$, again for the $N$=16 and $N$=50 cases. While the $d=0$ on-site correlations remain unchanged, the non-local correlations differ drastically between the two cluster sizes. Regardless of the physical or computation source of this cluster size discrepancy the examination of the spin-spin correlations gives an excellent metric to determine if sufficiently large clusters have been included and allows us to overcome this issue at low temperature by extrapolating only with clusters large enough to include all relevant correlations. For the data presented in this work, this will manifest as a natural minimum accessible temperature based on our maximum cluster size of 50 sites. This minima can be overcome by extrapolating with larger cluster sizes, but the precise clusters required become a detail of the observed spin correlations as in Fig.~\ref{fig:sisj}. More importantly, from this, one can see to what level spin-spin correlations are maintained in various cluster sizes. %As such, the omission of smaller clusters in extrapolations is warranted at low temperature as those results would not be expected to land precisely on the finite-size scaling curve. These are omitted if doing so would reduce the uncertainty in the extrapolation of the remaining larger clusters. In the infinite $U$ limit a system should contain no double occupancy at half-filling but for any finite $U$ this is not the case. In Fig.~\ref{fig:D}(a) we show the double occupancy obtained from clusters of size $N=$20, 50 and in the extrapolation to the TL. At very low temperature we see in the case of $N$=20 that the expected reduction in double occupancy for reduced temperature begins to reverse below $T\approx0.5$.\cite{paiva:2010} We also note (see inset) that as we push towards $N$=50 and the TL that the double occupancy at low temperature increases further. The rise in double occupancy, which is related to the potential energy, coincides with a continually decreasing total energy shown in the inset of Fig.~\ref{fig:D}(b). This indicates a reduction in kinetic energy which allows us to understand the rise in double occupancy as a physical consequence of the electrons becoming localized. This same effect has been phrased previously as a consequence of a low temperature increase in the local spin moment, $\langle {S_i^z}^2\rangle$ for reduced temperatures caused by a rise in double occupancy.\cite{paiva:2001} Examining higher temperature there is a behavioural shift in the double occupancy. This occurs in the range $T/t=1.0 \to 2.0$ where the double occupancy changes from the roughly constant value of $D=0.05$ to having a continued increase with temperature. With this in mind we can examine the low temperature behaviour of the energy in Fig.~\ref{fig:D}(b). At high temperatures we see a rise in energy which mimics the rise in $D$ above $T/t=2.0$. At low temperatures we see the need for large cluster sizes. For the smaller cluster of $N=20$ the energy is nearly smooth to temperatures as low as $0.1t$. We see however a shift which occurs only at low temperatures in the large clusters. This shift occurs at a temperature which reasonably agrees with the previously identified pseudogap temperature scale, $T^*\approx 0.3t$,\cite{paiva:2010} the temperature below which a reduction of the density of states is observed to occur in the antinodal direction but not in the nodal direction.\cite{macridin:2006, paiva:2001} While this feature is present in the 20-site case it is only extremely weak and, with the exception of Ref.~\onlinecite{mikelsons:2009}, has been mostly unmentioned in previous works which considered only smaller clusters. We note the agreement of our results with extrapolated NLCE data\cite{ehsan:priv} shown for intermediate temperatures in the inset of Fig.~\ref{fig:D}. %The extrapolated DCA results are in agreement with the NLCE data shown above our target minimum temperature of $T/t=0.2$. This comparison serves two purposes. First, within our uncertainty it verifies that the resummation scheme employed in NLCE is an accurate representation of the thermodynamic limit at temperatures below where the raw data diverges. Second, it raises a question regarding the true low temperature behaviour in the thermodynamic limit, at $T<0.2t$ where the extrapolated NLCE data begins to diverge and the DCA TL extrapolation begins to fail for the cluster sizes employed here. This can in principle be resolved using DCA, but in this work we cannot reasonably extend to lower temperatures, as the correlations (emphasized in Fig.~\ref{fig:sisj}) will require larger cluster sizes and this compounds the problem of computing at low temperature and accurately extrapolating DCA results to the TL. We also examine the spin correlations, $\langle S_i^z S_j^z \rangle_{nn} $ over the set of nearest-neighbours (nn) with variation in temperature plotted in Fig.~\ref{fig:D}(c). %There is a strong behavioural shift in the nearest-neighbour spin correlations at low temperature for different cluster size. Evidently, as is the case in Fig.~\ref{fig:D}(a), there is an underestimate of the double occupancy which results in the average spin correlation being skewed towards a larger negative value, suggesting a tendency towards antiferromagnetism at low temperature which is stronger in smaller sized clusters, in agreement with the behaviour of Fig.~\ref{fig:sisj} at low temperature. However, in this case we have omitted the data points at the lowest temperatures since there the uncertainties in this quantity become too large for a reliable extrapolation. \begin{figure} \begin{center} \includegraphics[width=0.95\linewidth]{fig3-4.eps} \end{center} \caption{\label{fig:esc}(Color online) Energy, $E(T)$, and entropy, $S(T)$, as functions of $T/t$ at half-filling extrapolated from DCA data to the TL for $U/t$=4, 8 and 12.} \end{figure} % NLCE data taken from Ref.~\onlinecite{khatami:2011}.\cite{ehsan:priv} In Fig.~\ref{fig:esc} we present results of the energies from DCA extrapolated to the TL for varied interaction strength at half-filling. As is apparent in other works,\cite{Parcollet04,gull:2008b,gull:2009,khatami:2011} the $U=4$ case does not show an incompressible phase at these temperatures. It is expected that the incompressible regime will have some impact on the intermediate and strongly coupled energies. While the effect is subtle in the energy, the cumulative effect on the entropy, shown in Fig.~\ref{fig:esc}(b), results in a decrease in $S(T)$ below $T^*$. On physical grounds this represents the loss of available thermal configurations at finite temperature as the electronic density of states becomes gapped and enters a partially gapped pseudogap state. This momentum-selective Mott transition is the same physics which explains the partially incompressible region of densities near half-filling at, for example, $T=0.25$ in Fig.~\ref{fig:nmu} and may may have consequences for the interplay between superconductivity and the pseudogap.\cite{gull:2012,gull:2009, werner:2009, gull:2010} Though such a depression exists in the strong coupling case for N=50 we cannot accurately extrapolate to the TL below this temperature with our current range of cluster sizes and limit our present work to $T>0.3t$ at the value of $U/t=12$. %{\color{red} Intermediate and strong coupling cases show a similar depression in the energies as one reduces the temperature into this incompressible regime. %The existence of this feature at $T^*$ is a possible indicator the pseudogap and its impact on the energetics of the electrons and may have consequences for the interplay between superconductivity and the pseudogap.\cite{gull:2012} %While this is difficult to see in the energies on this scale, the cumulative effect on the entropy, shown in Fig.~\ref{fig:esc}(b), results in a decrease in $S(T)$ below $T^*$. On physical grounds this represents the loss of available thermal configurations at finite temperature as the electronic density of states becomes gapped and enters a partially gapped pseudogap state. This momentum-selective Mott transition is the same physics which explains the partially incompressible region of densities near half-filling at, for example, $T=0.25$ in Fig.~\ref{fig:nmu}.\cite{gull:2009, werner:2009, gull:2010} %} %Overlaid with $S(T)$ is the NLCE data of Ref.~\onlinecite{khatami:2011} which has access to lower temperatures as $U/t$ is increased before diverging. In all cases, the DCA precisely reproduces the NLCE over the entire range of available temperatures. For the weaker coupling case of $U/t=4$ we continue to low temperature, where the NLCE cannot access, below which $S(T)\to 0$ and $E(T)$ becomes constant. In Fig.~\ref{fig:u8filled}(a) and \ref{fig:u8filled}(b) we present for $U/t=8$ the energy and entropy respectively for doping values near to but away from half-filling. In addition, the energies also provide direct access to the electronic specific heat shown in Fig.~\ref{fig:u8filled}(c). Our $C(T)$ data are obtained by taking finite differences in the spline interpolation of neighbouring energy values and therefore amplifies the numerical noise of Fig.~\ref{fig:u8filled}(a). For $C(T)$ we omit error bars as the value and uncertainty are somewhat dependent upon the method of interpolation and differentiation. Despite this, our results agree with the extrapolated NLCE data at half filling.\cite{khatami:2012} We have also extended the present work to include three dopings, of $n=0.85$, 0.90 and 0.95, away from half filling in the region most difficult for DCA calculations due to the occurrence of a sign problem. Here finite size issues in DCA result in deviations from standard DMFT results. Though not explored here, this present work shows that coarsely gridded and interpolated DCA data can be used to obtain precise specific heat data at dopings far away from half-filling, where other techniques cannot converge at low temperatures. Other Monte-Carlo works \cite{duffy:1997,paiva:2010, fye:1987} obtained on finite systems have identified the two main features of the specific heat, namely the low temperature spin and high temperature charge peaks. Here we present accurate results of the high temperature charge peak (near $T/t=2.0$) in the TL. At low temperatures we simply remark that the impact of the shift in energy, which is only apparent for large clusters, acts to create the spin peak in $C(T)$.\cite{khatami:2012} For cold-atom experiments both of these peaks in $C(T)$ will act as a strong barriers to further cooling of an atomic gas system. %It is difficult to comment on the existence of this low temperature `kink' away from half filling due to the temperature restriction imposed by the sign problem. However, one would expect that the creation of a pseudogap in the system's energy dispersion acts to reduce the density of states near the Fermi level. At finite temperature this should act to reduce the number of available thermal configurations %If it is indeed the case that this drop in The results presented in Figs.~\ref{fig:extrap} to \ref{fig:u8filled} include only a small part of the numerical results which we make available in this paper. For the sake of brevity we organize these additional results in the supplementary material which contains a detailed explanation of the data sets. In addition to the $U/t=8$ data we have presented here, we also include in the supplement the extrapolations to the thermodynamic limit for $U/t=4$ and 12 both at and away from half-filling. We expect these results to be a useful reference for comparison with other techniques in parts of phase space (in particular at low $T$, away from half-filling) where no previous controlled Monte-Carlo results exist. \begin{figure} \begin{center} \includegraphics[width=\linewidth]{fig-filled-4.eps} \end{center} \caption{\label{fig:u8filled}(Color online) Energy, $E(T)$, entropy, $S(T)$, and specific heat capacity, $C(T)$, as functions of $T/t$ extrapolated to the TL for $U/t=8$ for filling values of $n=0.85$, 0.9, 0.95, and 1.0 (half-filled). The extrapolated NLCE data in (c) can be found in Ref.~\onlinecite{khatami:2012}.} \end{figure} \section{Summary and Conclusions} \label{sec:conclusions} We have calculated the full thermodynamics of the 2D-Hubbard model by extrapolating DCA results on large clusters to the thermodynamic limit. Our results are numerically exact and, at high temperature, are validated against numerical linked-cluster expansion results. We have extended our parameter range substantially beyond what was previously shown. We provide results in the thermodynamic limit, for lower temperatures as well as for a wide range of filling values. We assert that our results are numerically exact within the errors we provide, verified by explicitly examining the range of spin correlations in real space. From this we can observe that our choice of cluster sizes has included all correlations. We note the occurrence of low temperature features in energy and entropy which seem to correlate with the onset of pseudogap physics at $T^*$ which are not captured directly in thermodynamic quantities for small clusters. Finally, we present exact results for nearest-neighbour spin correlations. Since $\langle S_i^z S_j^z\rangle_{nn}$ is measurable in cold-atom experiments, it may be used for thermometry.\cite{greif:2011,kollath:2006} Accurate values and reliable error bars are essential for this purpose. We have shown that DCA is an ideal technique for establishing the temperature dependence of these correlations, and have provided tables in the supplement which contain reference data needed for alternate techniques. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{acknowledgments} We thank Ehsan Khatami for useful discussions and the extrapolated NLCE reference data we present in this paper, and Richard Scalettar for investigating the discrepancy with Ref.~\onlinecite{paiva:2010} and providing us with updated DQMC data.\cite{scalettar:priv} Our continuous-time QMC codes are based on the ALPS libraries.\cite{gull:2011:alps, bauer:2011} \end{acknowledgments} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\end{acknowledgments} \bibliographystyle{apsrev4-1} \bibliography{bib}% Produces the bibliography via BibTeX. \end{document} }

66 — 1306.0735

\caption{Absolute bias (left panel) and variance (right panel) score estimates for $\tau$ from the autoregressive model using our $\mathcal{O}(N)$ algorithm with $\lambda = 0.99$ (\color{gray}$\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\ast$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\ast$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ \color{black}), $\lambda = 0.95$ (\color{orange}\protect\rule[0.5ex]{0.8cm}{0.7pt} \color{black}), $\lambda = 0.9$ (\color{green} $\boldsymbol{-} \cdot \times \boldsymbol{-} \cdot \times \boldsymbol{-}$ \color{black}), $\lambda = 0.8$ (\color{blue} $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ \color{black}), $\lambda = 0.7$ ($\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$), Fixed-lag smoother $L = 10$ (\color{red} $\cdot\triangledown\cdot\cdot\triangledown\cdot\cdot\triangledown\cdot$ \color{black}), and the Poyiadjis $\mathcal{O}(N)$ algorithm ($\boldsymbol{\cdot\cdot\cdot\cdot\cdot\cdot}$) and $\mathcal{O}(N^2)$ with $N = 500$ ($\boldsymbol{-} \cdot \otimes \boldsymbol{-} \cdot \otimes \boldsymbol{-}$).}

\caption{Absolute bias (left panel) and variance (right panel) of the $\phi$ component of the observed information matrix from the autoregressive model using our $\mathcal{O}(N)$ algorithm with $\lambda = 0.99$ (\color{gray} $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\ast$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\ast$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ \color{black}), $\lambda = 0.95$ (\color{orange} \protect\rule[0.5ex]{0.8cm}{0.7pt} \color{black}), $\lambda = 0.9$ (\color{green} $\boldsymbol{-} \cdot \times \boldsymbol{-} \cdot \times \boldsymbol{-}$ \color{black}), $\lambda = 0.8$ (\color{blue} $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ \color{black}), $\lambda = 0.7$ ($\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$), Fixed-lag smoother $L = 7$ (\color{red} $\cdot\triangledown\cdot\cdot\triangledown\cdot\cdot\triangledown\cdot$ \color{black}), and the Poyiadjis $\mathcal{O}(N)$ algorithm ($\boldsymbol{\cdot\cdot\cdot\cdot\cdot\cdot}$) and $\mathcal{O}(N^2)$ with $N = 500$ ($\boldsymbol{-} \cdot \otimes \boldsymbol{-} \cdot \otimes \boldsymbol{-}$).}

\caption{Root mean squared error of parameter estimates $\phi$ (left panel) and $\sigma$ (right panel) averaged over 20 Monte Carlo simulations from our $\mathcal{O}(N)$ algorithm with $\lambda=0.95$ (\protect\rule[0.5ex]{0.8cm}{0.6pt}), Poyiadjis $\mathcal{O}(N)$ ( \color{blue} $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\triangledown$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\triangledown$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ \color{black}), Poyiadjis $\mathcal{O}(N^2)$ (\color{green} $\boldsymbol{-} \cdot \Diamond \boldsymbol{-} \cdot\boldsymbol{-} \Diamond$ \color{black}), Fixed-lag smoother ( \color{cyan}$\protect\rule[0.5ex]{0.18cm}{0.9pt}$ $\circ$ $\protect\rule[0.5ex]{0.18cm}{0.9pt}$ $\protect\rule[0.5ex]{0.18cm}{0.9pt}$ $\circ$ $\protect\rule[0.5ex]{0.18cm}{0.9pt}$ \color{black}), Fixed-lag smoother ( \color{red} $\boldsymbol{\cdot}$ $\boldsymbol{\cdot}$ $+$ $\boldsymbol{\cdot}$ $\boldsymbol{\cdot}$ $+$ $\boldsymbol{\cdot}$ \color{black}) with score only and the Kalman filter estimate ($\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$).}

\caption{Batch (left panel) and recursive (right panel) parameter estimation for $\lambda=0.99$ (\protect\rule[0.5ex]{0.8cm}{0.6pt}), $\lambda=0.95$ ( \color{red}$\boldsymbol{-} \cdot \Diamond \boldsymbol{-} \cdot\boldsymbol{-} \Diamond$ \color{black}), $\lambda=0.5$ ( \color{cyan}$\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\times$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$ $\times$ $\protect\rule[0.5ex]{0.1cm}{0.5pt}$ \color{black}).}

\caption{Root mean squared error of parameter estimates $\phi$ (left panel) and $\sigma$ (right panel) averaged over 100 Monte Carlo simulations from our algorithm with $\lambda=0.95$ and $\phi = 0.9$ (\protect\rule[0.5ex]{0.8cm}{0.7pt}), $\phi=0.99$ (\color{cyan} $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.2cm}{0.7pt}$ \color{black}), $\phi=0.999$ (\color{green} $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\Diamond$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ \color{black}) and the particle learning algorithm with $\phi = 0.9$ (\color{blue} $\cdot\triangledown\cdot\cdot\triangledown\cdot\cdot\cdot$ \color{black}), $\phi = 0.99$ (\color{red} $\boldsymbol{-} \circ \cdot \boldsymbol{-} \circ \cdot\boldsymbol{-}$ \color{black}), $\phi = 0.999$ ($\boldsymbol{\cdot\cdot}\times\boldsymbol{\cdot\cdot}\times\boldsymbol{\cdot\cdot}$).}

\caption{Parameter estimates for $\mu_2$ (left panel) and $\mu_4$ (right panel) from our $\mathcal{O}(N)$ algorithm with $\lambda = 0.95$ (\color{orange}\protect\rule[0.5ex]{0.8cm}{0.7pt}\color{black}), $\lambda = 0.7$ ($\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$ $\protect\rule[0.5ex]{0.1cm}{0.7pt}$ $\vartriangle$ $\protect\rule[0.5ex]{0.4cm}{0.7pt}$), Fixed-lag smoother $L = 5$ (\color{red} $\cdot\triangledown\cdot\cdot\triangledown\cdot\cdot\triangledown\cdot$ \color{black}), and the Poyiadjis $\mathcal{O}(N)$ algorithm (\color{gray} $\boldsymbol{\cdot\cdot\cdot\cdot\cdot\cdot}$ \color{black}) and $\mathcal{O}(N^2)$ ($\boldsymbol{-} \cdot \otimes \boldsymbol{-} \cdot \otimes \boldsymbol{-}$).}

67 — 1306.2702

\caption{ %\color{Blue} Trajectories starting from initial states, $\bfv{u}_{(a,b)}$ for a different set of parameters $(a,b)$, separate towards either the laminar $\bfv{u}_{\mbox{\tiny L}}$ (grey dashed curves) or the turbulent states (black solid curves), which are obtained at $Re=1000$. %An initial state consists of $\bfv{u}_{\mbox{\tiny H}}$($\triangle$) with a perturbation, which does not satisfy the reflection symmetry in the case of $y>0$. In the case of $\eta>0$, the initial state does not satisfy the reflection symmetry. %The final state, either laminar or turbulent, is determined only by a subtle difference of magnitude of additive perturbation to the initial state. Some of trajectories starting around $\bfv{u}_{\mbox{\tiny H}}$ at $Re=1000$ ({\color{red} $\triangle$}) with a particular set of $(a,b)$ trace a part of the BB as ``watershed'', which divides the whole space into the basins of the turbulent and laminar attractors. This BB is outlined by a hetero-clinic orbit\cite{hal09} connecting $\bfv{u}_{\mbox{\tiny H}}$ and $\bfv{u}_{\mbox{\tiny N}}$ at $Re=1000$ ({\color{red} $\bigtriangledown$}). %, which is likely to be realized even at relatively high $Re$. For reference, $\bfv{u}_{\mbox{\tiny H}}$ and $\bfv{u}_{\mbox{\tiny N}}$ obtained at $Re=300, 10000$ are plotted as well (grey triangles), both of which approach towards $\xi=0$ with increasing $Re$. % All asymptotic behaviours indicate that the HVS instead of NBW can be referred to as the criterion of transition at high Reynolds numbers . }

68 — 1306.2831

\caption{\label{Fig:RMT:HPI:lambda:1234} ({\it{A}}) Evolution of the average correlation coefficient. The horizontal red line shows the critical value at significance level 5\% of the correlation coefficient at each time $t$. The error bar is the standard deviation of the PDF at each time $t$. For the evolution of the PDF, see {\color{blue}{Fig.~S1}}. ({\it{B}}) Evolution of the five largest eigenvalues $\lambda_n$ of $\mathbf{C}(t)$ with $n=1,2,3,4$, and $5$. The horizontal dot-dashed red line is the maximum eigenvalue $\lambda_{\max}$ predicted by the RMT and the horizontal red line represents the critical values $\lambda_{5\%}$ at the significance level of 5\%. The five vertical dashed lines corresponding to the five regime-shift points. ({\it{C}}) Evolution of absorption ratio $E_n(t)$ for $n=1,2,3,4$, and $5$.}

\caption{\label{Fig:RMT:HPI:MarketEffect} Market effect hidden in the largest eigenvalues. ({\em{A}} to {\em{D}}) Evolution of the correlation coefficient $k_n(t)$ between $R_n$ and $R$ in each moving window. The blue symbols are estimated using ordinary least-squares linear regression, while the red ones are estimated using robust fitting. The four vertical lines indicate four regime-shift points ${\mathcal{T}}_1$ between $1993Q3$ and 1993Q4, ${\mathcal{T}}_2$ between $1997Q1$ and 1997Q2, ${\mathcal{T}}_3$ between $1999Q2$ and 1999Q3, and ${\mathcal{T}}_4$ between $2002Q2$ and 2002Q3, separating five different regimes. The shading area in each plot means that the associated eigenvalue contains a market effect in the corresponding time period. See {\color{blue}{Fig.~S3}} for the scatter plots of $R_n$ against $R$. }

69 — 1306.4012

\caption{\label{figura4}}{\includegraphics[scale=0.25]{figura4.eps}}

70 — 1306.4153

\caption{The fit values are the best fit marginalised values with \Mnu{} set to the 95\% upper limit. \red{NOT SURE TABLE NEEDS TO BE IN PAPER, just a reminder for me}}

71 — 1306.5263

\caption{Notation for the quantities in our formulation. % Quantities in black are provided. % Quantities in \red{red} are hidden. % Quantities in \green{green} are learned.}

72 — 1306.5783

\caption{(Color online) Frequency shift as a function of tip angle, $\beta$, at $\frac{T}{T_c} \approx 0.86$ with the strain axis perpendicular to the magnetic field. The black solid curve is the expected behavior for either easy-plane or easy-axis models, \red{and consequently} for this orientation of the magnetic field it is not possible to distinguish between them. }

\caption{(Color online) Frequency shift as a function of temperature for different aerogel orientations. For strain axis perpendicular to the magnetic field, \red{the} frequency shift with $\beta=8^{\circ}$ is shown as the blue circles. For strain axis parallel to the magnetic field, the frequency shift with $\beta=16^{\circ}$ and $51^{\circ}$ are shown as red squares and green triangles respectively. The frequency shift with $\beta=51^{\circ}$ (green triangles) was scaled to the $\beta=0$ limit, using Eq.~1, to better compare with other data. The black solid (black dashed) \red{curves indicate} the maximum positive (negative) frequency shift \red{in the} small tip angle limit, Eq.~2 (Eq.~1).}

73 — 1306.5935

\caption{Energy density plots of the multi-soliton solutions for $N=5$ and $B\leq4$ (colouring is based on the segment in which the point lies in the target space). Note that the \protect\includegraphics[scale=0.12,natwidth=121,natheight=82]{Images/forms/pent-4-7.jpeg} solution was not obtained, although we still expect this solution to exist. It was very similar to the \protect\includegraphics[scale=0.12,natwidth=90,natheight=80]{Images/forms/B2.jpeg} caveat in (m) meaning it was difficult to pick out initial conditions that would relax to the desired solution rather than this lower energy caveat form.}

74 — 1306.6141

\caption{Effect of $P_{e,k}$ on $g_{k}(\tau_{k})$ when $p_{w_{k}}(\tau_{k})=\frac{\epsilon}{2\alpha\Gamma(1/\epsilon)}\exp\left[-\left(\frac{|\tau_{k}|}{\alpha}\right)^{\epsilon}\right]$; $\alpha=1$, $\epsilon\in\{3,4\}$ and $P_{e,k}\in\{0,0.1,0.2\}$.\label{fig:bimodality_gtau}} \end{figure} Substituting $\tau_{k}=0$, $k\in\mathcal{K},$ in Eq.~(\ref{eq:Rao Test explicit- BSC case}), leads to the following simplified expression for threshold-optimized Rao test (denoted with $\Lambda_{\mathrm{R}}^{*}$): \begin{equation} \Lambda_{\mathrm{R}}^{*}=\frac{4\cdot\left[\sum_{k=1}^{K}(1-2\, P_{e,k})\cdot p_{w_{k}}(0)\cdot h_{k}\cdot(y_{k}-\frac{1}{2})\right]^{2}}{\sum_{k=1}^{K}(1-2\, P_{e,k})^{2}\cdot p_{w_{k}}^{2}(0)\cdot h_{k}^{2}}\label{eq:Rao_test_simplified_error_prone} \end{equation} which is considerably simpler than the GLRT, as it obviates solution of an optimization problem (which depends on $p_{w_{k}}(\cdot)$). Furthermore, the corresponding optimized non-centrality parameter, denoted with $\lambda_{Q}^{*}$, is given by: \begin{equation} \lambda_{Q}^{*}=4\,\theta^{2}\cdot\sum_{k=1}^{K}\left[(1-2\, P_{e,k})^{2}\cdot p_{w_{k}}^{2}(0)\cdot h_{k}^{2}\right]\label{eq:optimized-noncentralitypar-BSC-1} \end{equation} \emph{Remarks} - In the case of BSCs of the same quality (i.e. $P_{e,k}=P_{e},$ $k\in\mathcal{K}$) we simply get $\lambda_{Q}^{*}=(1-2\, P_{e})^{2}\cdot\lambda_{Q_{0}}^{*},$ where $\lambda_{Q_{0}}^{*}\triangleq4\,\theta^{2}\cdot\sum_{k=1}^{K}\left[p_{w_{k}}^{2}(0)\cdot h_{k}^{2}\right]$ represents $\lambda_{Q}^{*}$ in the ideal BSC case ($P_{e,k}=0$, $k\in\mathcal{K}$). This result generalizes the one in \cite{Fang2013}, by stating that the \emph{loss due to non-ideal communications is asymptotically independent} \emph{of} $p_{w_{k}}(\cdot)$, $k\in\mathcal{K}$. \section{Comparison in homogeneous scenario\label{sec:GLRT and Rao - homogeneous scenario}} In this section we study the simplified scenario $h_{k}=h$, $p_{w_{k}}(\cdot)=p_{w}(\cdot)$, $P_{e,k}=P_{e}$, $k\in\mathcal{K}$, to get an intuitive interpretation of the two threshold-optimized tests ($\tau_{k}^{*}=0$). Based on these assumptions, the statistics in Eqs. (\ref{eq:GLRT_general}) and (\ref{eq:Rao_test_simplified_error_prone}) reduce to: \begin{eqnarray} \Lambda_{\mathrm{G}}^{*} & = & 2K\cdot\left[\hat{\rho}\ln\left(\frac{\hat{\rho}}{\rho_{0}}\right)+(1-\hat{\rho})\ln\left(\frac{1-\hat{\rho}}{1-\rho_{0}}\right)\right]\label{eq:GLRT_KL_pre}\\ & = & 2K\cdot D_{\mathrm{KL}}(\hat{P}(y_{k})\parallel P(y_{k};\theta_{0}))\label{eq:GLRT_hoeffding_KL}\\ \Lambda_{\mathrm{R}}^{*} & = & 4K\cdot\left[\hat{\rho}-\rho_{0}\right]^{2}\label{eq:Rao_TVD_pre}\\ & = & 4K\cdot\left[D_{\mathrm{TVD}}(\hat{P}(y_{k})\parallel P(y_{k};\theta_{0}))\right]^{2}\label{eq:Rao_TVD} \end{eqnarray} where $\Lambda_{\mathrm{G}}^{*}\triangleq\Lambda_{\mathrm{G}}(\tau_{k}=0)$, $\hat{\rho}\triangleq\sum_{k=1}^{K}y_{k}/K$ and $\rho_{0}\triangleq\nicefrac{1}{2}$. Here $\hat{P}(y_{k})$ represents the empirical distribution of the i.i.d. binary source $\{y_{1},\ldots,y_{K}\}$ and $D_{\mathrm{KL}}(\cdot\parallel\cdot)$ and $D_{\mathrm{TVD}}(\cdot\parallel\cdot)$ denote the \emph{Kullback-Leibler }(KL) and \emph{total variation distance }(TVD) divergences, respectively \cite{Cover2006}. It is worth noticing that in Eq. (\ref{eq:GLRT_hoeffding_KL}) we exploited the closed form of $\hat{\theta}_{1}=-\frac{1}{h}F_{w}^{-1}\left(\left(\hat{\rho}-P_{e}\right)/\left(1-2P_{e}\right)\right)$ (see \cite{Ribeiro2006a} for a similar result). Exploiting KL% \footnote{Since it is increasing when $\hat{\rho}>\rho_{0}$ and symmetric around $\rho_{0}$.% } and TVD divergences properties it can be shown that both Eqs. (\ref{eq:GLRT_hoeffding_KL}) and (\ref{eq:Rao_TVD}) are monotone (increasing) functions of $|\hat{\rho}-\rho_{0}|$ and therefore \emph{represent equivalent tests} in a homogeneous sensor scenario, meaning their performances coincide also for a finite number of sensors. Finally, we derive a tighter asymptotic form of the conditional pdf (\emph{not requiring} the weak-signal assumption) of both the tests in this scenario with the help of the central limit theorem (CLT) \cite{Cover2006}. Without loss of generality we focus hereinafter on $\Lambda_{\mathrm{R}}^{*}$ (since $\Lambda_{\mathrm{G}}^{*}$ has the same performance). For this purpose, we define the RV $\xi\triangleq\frac{\sum_{k=1}^{K}\left(2y_{k}-1\right)}{\sqrt{K}}$ and we consider the asymptotic form of $p_{\xi}(\cdot|\mathcal{H}_{i})$, $i\in\{0,1\}$, which according to the CLT is given as $K\rightarrow+\infty$ by: \begin{gather} \xi|\mathcal{H}_{0}\;\overset{{\scriptstyle a}}{\sim}\;\mathcal{N}(0,1)\qquad\quad\xi|\mathcal{H}_{1}\;\overset{{\scriptstyle a}}{\sim}\;\mathcal{N}(\sqrt{K}\tilde{\mu}_{1},\tilde{\sigma}_{1}^{2})\label{eq:CLT-based performance} \end{gather} where $\tilde{\mu}_{1}\triangleq(1-2P_{e})(2\rho_{1}-1)$ , $\tilde{\sigma}_{1}^{2}\triangleq4\cdot[1+P_{e}(2\rho_{1}-1)-\rho_{1}]\cdot[\rho_{1}+(1-2\rho_{1})P_{e}]$ and $\rho_{1}\triangleq F_{w}(-h\theta)$. From inspection of Eq.~(\ref{eq:Rao_TVD_pre}), it can be readily verified that $\Lambda_{\mathrm{R}}^{*}=\xi^{2}$ holds, which can be exploited to obtain closed form performance expressions. \section{Numerical Results\label{sec:Numerical-Results}} In this section we compare the Rao test to the GLRT. We evaluate the performance in terms of system false alarm and detection probabilities, defined as $P_{F_{0}}\triangleq\Pr\{\Lambda>\gamma|\mathcal{H}_{0}\}$ and $P_{D_{0}}\triangleq\Pr\{\Lambda>\gamma|\mathcal{H}_{1}\}$, respectively, where $\Lambda$ is the statistic employed at the DFC. We also define the $k$th sensor observation signal-to-noise ratio (SNR) as $\Gamma_{k}\triangleq\left(h_{k}^{2}\theta^{2}/\mathbb{E}\{w_{k}^{2}\}\right)$. In Fig. \ref{fig: ROC} we illustrate $P_{D_{0}}$ vs $P_{F_{0}}$ in a WSN with $K=5$ sensors where $\theta=1$, $h_{k}\sim U(0,a)$, $k\in\mathcal{K}$ (but known at the DFC), and two noise pdfs: ($i$) $w_{k}\sim\mathcal{N}(0,\sigma_{k}^{2})$ and $(ii)$ $w_{k}\sim\mathcal{L}(0,\beta_{k})$, such that $\mathbb{E}\{w_{k}^{2}\}=1$. We consider four combinations corresponding to $P_{e,k}=P_{e}\in\{0,0.2\}$ and $\bar{\Gamma}_{dB}\in\{0,10\}$, where we have denoted $\bar{\Gamma}\triangleq\mathbb{E}\{\Gamma_{k}\}$ (in our case $\bar{\Gamma}=(a^{2}\theta^{2})/3\cdot\mathbb{E}\{w_{k}^{2}\}$) as the average observation SNR. The figures are based on $10^{5}$ Monte Carlo runs. First, it is apparent that the performances of the GLR and the Rao tests are practically the same for all the considered scenarios; however the implementation of the Rao test is much simpler than that of the GLRT. Also, the difference in performances under Laplacian and Gaussian noises is significant only at $\bar{\Gamma}_{dB}=0$, while at $\bar{\Gamma}_{dB}=10$ the curves almost overlap. This is explained since when $\bar{\Gamma}$ is low the signal is more concentrated around zero. Then the imbalance in the binary pmf observed at the output of each quantizer is higher when $w_{k}\sim\mathcal{L}(0,\beta_{k})$. In Fig. \ref{fig: Pd0 vs K} we show $P_{D_{0}}$ as a function of $K$, assuming $P_{F_{0}}=0.1$. We consider $\theta=0.5$, $h_{k}=1$ and two noise pdfs: ($i$) $w_{k}\sim\mathcal{N}(0,\sigma_{k}^{2})$ and $(ii)$ $w_{k}\sim\mathcal{L}(0,\beta_{k})$, such that $\mathbb{E}\{w_{k}^{2}\}=1$ (thus $(\Gamma_{k})_{dB}\approx-6$), $k\in\mathcal{K}$. Also, we consider $P_{e}\in\{0,0.2\}$, thus determining a homogeneous scenario. First, Monte Carlo simulations confirm the theoretical coincidence between the Rao test (bullet markers) and the GLRT (square markers). Secondly, it is apparent that the CLT-based performance expressions (dash-dot) are as accurate as those based on the weak-signal assumption (solid lines) for Gaussian noise, while in the Laplacian Case the weak-signal distribution is far from being representative of the distribution. Interestingly, when $P_{e}=0$, $\lambda_{Q_{0}}^{*}$ in the Laplacian case coincides with the non-centrality parameter achieved by a GLRT (or Rao test) based on the raw $x_{k}$, $k\in\mathcal{K}$, given by $\lambda_{UQ}=\theta^{2}\sum_{k=1}^{K}\frac{h_{k}^{2}}{\beta_{k}^{2}}$, that is Eq. (\ref{eq:optimized-noncentralitypar-BSC-1}) \emph{does not predict the loss due to quantization}. On the other hand, by exploiting the CLT-based performance in Eq. (\ref{eq:CLT-based performance}), we can compare $\lambda_{UQ}$ with the modified deflection coefficient of the asymptotic problem given by Eq. (\ref{eq:CLT-based performance}) $d_{Q}\triangleq K\tilde{\mu}_{1}^{2}/\tilde{\sigma}_{1}^{2}$, which for the Laplacian noise is given by $d_{Q}=\frac{K\theta^{2}(h^{2}/\beta^{2})}{[1-\exp(-|h\theta|/\beta)]^{2}}\cdot[1-(1-\exp(-\frac{|h\theta|}{\beta})^{2}]$; thus for this problem we have $\lambda_{UQ}/d_{Q}\approx1.45$, which predicts the performance loss well. \begin{figure}[t] \centering{}\includegraphics[width=0.94\columnwidth]{ROC_K5_rev_split}\caption{$P_{D_{0}}$ vs $P_{F_{0}}$; WSN with $K=5$ sensors, $h_{k}\sim\mathcal{U}(0,a)$, $\theta=1$, $\mathbb{E}\{w_{k}^{2}\}=1$ for Gaussian and Laplace noise; $P_{e}\in\{0,0.2\}$, $\bar{\Gamma}_{dB}\in\{0,10\}$.\label{fig: ROC}} \end{figure} \begin{figure}[t] \centering{}\includegraphics[width=0.88\columnwidth]{PdvsK_rev}\caption{$P_{D_{0}}$ vs $K$; $P_{F_{0}}=0.1$. Setup: $\theta=0.5$; $h_{k}=1$, $\mathbb{E}\{w_{k}^{2}\}=1$ ($(\Gamma_{k})_{dB}\approx-6$), $P_{e,k}\in\{0,0.2\}$, $k\in\mathcal{K}$ (homogeneous scenario). Square ($\square$) and bullet ($\bullet)$ markers refer to GLRT and Rao test, respectively; solid and dash-dot lines refer to weak-signal and CLT-based asymptotic pdfs, respectively. \label{fig: Pd0 vs K}} \end{figure} \section{Conclusions\label{sec:Conclusions}} We studied the Rao test for decentralized detection with an unknown deterministic signal as an attractive alternative to GLRT for a general model with quantized measurements, zero-mean, unimodal and symmetric noise (pdf), non-ideal and non-identical BSCs. The asymptotically optimal sensor thresholds were shown to be zero for many pdfs of interest and a fair choice in other scenarios; this result was exploited to simplify further the Rao test formula. Furthermore, it was shown through simulations that the Rao test, in addition to being asymptotically equivalent to the GLRT, achieves practically the same performance in the finite number of sensors case; for the case of homogeneous sensors a theoretical coincidence of the two tests was established. In such a scenario a general asymptotic performance were derived based on the CLT and not requiring the weak-signal assumption. These latter were shown to be crucial in performance analysis with peaked noise pdfs. \bibliographystyle{IEEEtran} \bibliography{IEEEabrv,sensor_networks} \end{document} }}

75 — 1306.6236

\caption{The magnetic moment of \lacatbcodve\measured under various fields in the FC ($\square$) and ZFC(\textcolor{red}{$\bigcirc$}) regimes. The insets show the corresponding inverse susceptibility graphs.}

76 — 1306.6407

\caption{Distribution of the \red{$K$-band} limiting magnitude $\mlim$ of the UKIDSS galaxies sample as a function of AGN redshift. Each point denote $\mlim$ for the area around each AGN sample. 11,335 AGN samples of $z=0.1$ -- 1.0 are distributed in the survey area of UKIDSS LAS. }

\caption{Distribution of the \red{$K$-band} absolute limiting magnitude $M_{\rm limit} = \mlim - DM$ (red crosses) and absolute threshold magnitude $M_{\rm th} = \mth - DM$ (green circles) of the UKIDSS galaxies around each AGN as a function of AGN redshift. Where $DM$ is the distance modulus. $m_{\rm th}$ is a threshold magnitude defined in Equation (\ref{eq:det_eff}), above which detection efficiency for galaxies becomes lower than 1.0. AGN samples are the same as Figure~\ref{fig:mlimit-z}. }

\caption{Distribution of the average number density $\rho_{0}$ of detectable galaxies ($<m_{\rm limit}$, red crosses) and galaxies brighter than $m_{\rm th}$ (blue circles) at the AGN redshift for each AGN sample. %We use samples of $\rho_{0} > 10^{-4}$ in the following analysis. \red{Among the 11,335 AGN samples plotted in this figure, 10,482 AGNs with $\rho_{0} > 10^{-4}{\rm Mpc}^{-3}$ for detectable galaxies (red crosses above the dashed line) are use in the following analysis. For the analysis of the complete galaxy sample in Section \ref{completeS}, we use 6,107 AGNs with $\rho_{0} > 10^{-4}{\rm Mpc}^{-3}$ for galaxies brighter than $m_{\rm th}$ (blue circles above the dashed line). } }

\caption{Distribution of the surface number density of galaxies around AGNs as a function of the \red{$K$-band} limiting magnitude. $n_{7 \rm Mpc}$ is the surface density of UKIDSS galaxies in the circular area around the AGN within an angular radius corresponding to 7Mpc(comoving) at the AGN redshift. We select AGN samples whose deviations of $\log n_{7 \rm Mpc}$ are less than $1.5\sigma$ (between the two solid lines) to reject effect of foreground clusters and bad quality regions of UKIDSS data. AGN samples are the same as Figure~\ref{fig:mlimit-z}. }

\caption{Parameters of the Schechter function derived by \red{ Cirasuolo et al. 2007 ($K$-band. $z\geq0.5$), Gabasch et al. 2004 (150nm, 280nm, $u'$, and $g'$. $z\geq0.6$) Gabasch et al. 2006 ($r'$, $i'$, and $z'$. $z\geq0.6$) Kochanek et al. 2001 ($K$-band. $z=0$) Montero-Dorta $\&$ Prada 2009 ($u'$, $g'$, $r'$, $i'$, and $z'$. $z=0.1$). } The solid lines represent fitting functions to parametrize $M_{*}$ and $\Phi(M_{0})$ as a function redshift. {\it Top panel}: $M_{*}$ for each rest-frame wavelength band. {\it Bottom panel}: Number densities $\Phi(M_{0})$ at a reference magnitude $M_{0}$. The reference magnitude $M_{0}$ for each wavelength band is $-18$ for 150~nm, 280~nm, and $u'$ band, $-20$ for $g'$ band, $-21$ for $r'$, $i'$, and $z'$ band, and $-22$ for $K$ band. }

77 — 1306.6857

\caption{Simulation results in the carbon wall and the ILW under the identical conditions are presented: (a)Plasma current, (b)electron density(blue) and deuterium atom density(red), (c)electron temperature(blue) and ion temperature(red), (d)radiation power losses \textcolor{Correction}{and ohmic heating power}(dashed black: total radiation power loss, solid blue: deuterium radiation, solid red: beryllium radiation, solid black: carbon radiation, dashed blue: oxygen radiation, \textcolor{Correction}{and dashed red: ohmic heating power}), (e)sputtering yield(solid black: carbon sputtering due to incident deuterium ion, solid blue: beryllium sputtering due to incident deuterium ion, solid red: self-sputtering yield), (f)impurity densities in each charge state.}

78 — 1307.0081

\caption{\label{fig:gegm}Profiles of the normalised $\ExB$ shearing rate $\gEbar$ (solid line) calculated by TRANSP for the equilibrium from \#27268 at$\quant{0.25}{s}$ and the normalised linear growth rates $\gm/(\vti/a)$ calculated by GS2 for cases I (\textcolor{magenta}{$\triangle$} -- with adiabatic electrons, no flow) and II (\textcolor{blue}{$\diamond$} -- kinetic electrons, no flow or collisions) for the similar equilibrium from \#22807 at$\quant{0.25}{s}$. \color{black}}

79 — 1307.0563

\caption{Absorption-profile fits for the HVF-strong and slowly declining SN~2005cf (top row); the HVF-weak and rapidly declining SN~2004gs (middle row); and the HVF-weak, slowly declining, and high-velocity (HV) SN~2002cd (bottom row). Profiles have been normalised by the fitted pseudo-continuum, with data shown in blue and the full fitted profile in dashed red. For the \canir\and\cahk\profiles (middle and right columns, respectively), the fitted photospheric absorption component is indicated as the thin magenta curve, while the fitted HVF component is the thick cyan curve. For\cahk, the \siblue\component is shown as the dotted green curve. The absorption-weighted velocity for each line is indicated as the vertical dotted black line in each panel (see text for details).}

\caption{Top: HVF strength \rhvf\versus the$(B-V)$ colour at maximum light. Middle: Absorption-weighted \CaII\velocities from both the\cahk\(filled cyan circles) and \canir\(open red squares) versus colour. Bottom: Absorption equivalent width (pEW) of the HVF component of the \canir\versus colour.}

80 — 1307.2748

\caption{Two machine-system. Phase angles \textcolor{meinhellblau}{$\delta_1$} (solid), \textcolor{meinhellblau}{$\delta_2$} (dashed), phase difference \textcolor{meinmagenta}{$\delta_{12}$}, angular velocities \textcolor{meindunkelblau}{$\omega_1$} (solid), \textcolor{meindunkelblau}{$\omega_2$} (dashed), voltages \textcolor{meinorange}{$E_1$}, \textcolor{meinorange}{$E_2$} and power transfer \textcolor{meingruen}{$P_{12}$} from $M_1$ to $M_2$ as functions of time. $\gamma_1=\gamma_2=0.2$, $P_{\text m1}=-P_{\text m2}=0.5$, $\alpha_1=\alpha_2=2.0$, $E_{\text f,1}=E_{\text f,2}=1.0$, $X_1=X_2=1.0$, $B_{11}=B_{22}=-0.8$, $B_{12}=B_{21}=1.0$. $P_\text{dist}=1.0$ during $t\in[10,12]$ (disturbance period denoted by vertical dotted lines). Left: Extended model. Right: Classical model. The machine voltages \textcolor{meinorange}{$E_1$} and \textcolor{meinorange}{$E_2$} are congruent because of identical machine and line parameters. Both systems return to stationary operation.}

\caption{Two-machine system. Disturbance scenario with $P_\text{dist}=1.5$\,. The other parameter values as in fig.\,\ref{fig:two-machine_1}. Left: Extended model. \textcolor{meinhellblau}{$\delta_1$}, \textcolor{meinmagenta}{$\delta_{12}$} (\textcolor{meinhellblau}{$\delta_2$}) display unbounded growth (decrease) beyond the shown interval. Right: Classical model. Unlike the classical system, the extended system is not stable in the sense of power system stability.}

\caption{Two-machine system. Disturbance scenario with $P_\text{dist}=2.5$\,. The other parameter values as in fig.\,\ref{fig:two-machine_1}. Left: Extended model. \textcolor{meinhellblau}{$\delta_1$}, \textcolor{meinmagenta}{$\delta_{12}$} (\textcolor{meinhellblau}{$\delta_2$}) display unbounded growth (decrease) beyond the shown interval. Right: Classical model. Both systems are unstable in the sense of power system stability.}

\caption{Six-machine system of three generators \textcolor{blue}{G$_i$} ($i=$1,3,5) and three motors \textcolor{red}{M$_i$} ($i=$2,4,6). Shunt susceptances are neglected. The susceptance matrix $\{B_{ij}\}_{i,j=1,..,6}$ is calculated from the transfer susceptances (cf.\,\ref{sec:representation_powerflow}) $B_{\text T,12}$=\,-1.0,$B_{\text T,23}$=\,-0.5,$B_{\text T,34}$=\,-0.7,$B_{\text T,45}$=\,-1.0,$B_{\text T,56}$=\,-1.2,$B_{\text T,16}$=\,-0.8,$P_{\text m,1}$=0.25, $P_{\text m,2}$=\,-0.2,$P_{\text m,3}$=0.2, $P_{\text m,4}$=\,-1.5,$P_{\text m,5}$=1.5, $P_{\text m,6}$=\,-0.25,$\gamma_i$=0.1 $\forall i$, the other machine parameters as in fig.\,\ref{fig:two-machine_1}.}

\caption{Six-Machine system. Disturbance scenario with $P_\text{dist}=1$ for $t\in[10,13]$. Angular velocities \textcolor{meindunkelblau}{$\omega_i$}, voltages \textcolor{meinorange}{$E_i$} ($i=1$: solid line, $i=2,..,6$ dashed lines) and power transfer \textcolor{meingruen}{$P_{ij}$} along all links (the reference machines are chosen in such a way that all stationary power flows are positive) as functions of time $t$. Left: Extended model. Right: Classical model. Both systems return to stationary operation.}

81 — 1307.2915

\caption{The distribution of record processing times in the {\reducewrite} sub-phase of {\reduce} tasks (Job$_2$).}

\caption{The illustration of task, sub-task, and sub-phase in a \ha job: (a) a \map\task consists of\map\and\merge\sub-tasks; and (b) a\reduce\task consists of\shuffle, \sort, and \reduce\sub-tasks.}

82 — 1307.3631

\caption{(Color online) Calculated principal peak energy ({\color{red}$\circ$}) vs. the reciprocal of the effective Na cluster radius. The TDLDA results for three structurally unoptimized clusters (see Table I) are also shown ($-$). The results from previous LDA/electron spill-out model ({\color{blue}$\diamond$})~\cite{WIJ06} and TDLDA ({\color{olive}$\triangle$})~\cite{WIJ06} calculations based on a spherical jellium model, as well as the experimental data ({\color{red}$\oplus$}~\cite{SK91} and {\color{red}$\bullet$}~\cite{XYK09}), are also displayed for comparison. Dotted, dashed, and dash-dotted lines, which converge to $\omega_{Mie}$, are a guide to the eye only, for the open diamonds, open triangles, and all types of circles, respectively. While the coupling of the surface plasmon with electron-hole transitions leads to the red-shift (dotted to dashed line) of the principal peak energies~\cite{WIJ06}, the ionic effect is evidently responsible for the further red-shift (dashed to dash-dotted line) that brings our results in good agreement with the experiments.}

\caption{(Color online) Calculated static polarizability ({\color{red}$\circ$}) vs. the effective Na cluster radius, with a +20 \% bar accounting for possible temperature effect up to 500 K being added. The results ($\mathbf{-}$) obtained using Sternheimer equation within TDDFT formalism\cite{ABMR07} for some clusters are also plotted. For comparison, experimental data for closed-shell clusters ({\color{red}$\bullet$})~\cite{TK01} (with error bar present) and ({\color{red}$\oplus$})~\cite{RA99} are also plotted, together with previous measurements on non-closed-shell clusters ({\color{blue}$\times$})~\cite{TK01} and theoretical results for closed-shell ({\color{olive}$\blacktriangle$}) and non-closed-shell ({\color{olive}$\triangle$}) clusters from the TDLDA/jellium background calculations~\cite{Eka85}.}

\caption{(Color online) The widths (FWHMs) ({\color{red}$\circ$}) of the principal peaks derived from the absorption spectra from the present work (Fig. 2) vs. the effective Na cluster radius. Also plotted are the FWHMs from previous matrix RPA-LDA/jellium model calculations ({\color{olive}$\triangle$})~\cite{Yan93} and the experimental widths ({\color{red}$\bullet$}) of Na$_{20}$ and Na$_{92}$~\cite{XYK09} derived from the absorption spectra shown in Fig. 4.}

83 — 1307.3880

\caption{An example of analysis of partly cloudy night sky. The image on the left shows RAW image of the sky. The image on the right shows the analysis results, the yellow crosses indicate the detected stars, green circles catalogue stars and red circles represent catalogue stars without detected pairs - the region covered by cloudiness. The {\color{red} RED} circle indicates the limiting zenith angle ($60^{\circ}$). }

84 — 1307.4834

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$ (right) for Shift contamination, $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $d_x=2$, $\alpha=0.5$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$ (right) for Shift contamination, $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $d_x=8$, $\alpha=0.5$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$e (right) for Point-mass contamination, $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $d_x=2$, $\alpha=0.5$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$ (right) for Point-mass contamination, $\varepsilon=\{0.1,\ldots,0.4\}$, $p=\{4,\ldots,16\}$, $d_x=8$, $\alpha=0.5$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$ (right) for Shift and Point-mass contamination, $\varepsilon=\{0.1,0.2\}$, $p=\{4,\ldots,16\}$, $\alpha=0.75$, $d_x=2$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

\caption{$\mbox{Bias}(\hat{\pmb\theta})$ (left) and $\mbox{Mis.Rate}(I_c,H^+)$ (right) for Shift and Point-mass contamination, $\varepsilon=\{0.1,0.2\}$, $p=\{4,\ldots,16\}$, $\alpha=0.75$, $d_x=8$ as a function of $\nu$. \textcolor{fmcd}{\underline{FastLTS}}, \textcolor{fmve}{\underline{FastS}}, \textcolor{RCS}{\underline{FastRCS}}.}

85 — 1307.5113

\caption{Real parts of the frequency dependent conductivity for the Drude-Lorentz function (\dashed) for the normal state and the Mattis-Bardeen theory (\full) for a BCS superconductor using $\omega_g/2\pi=1$~THz and $2\pi/\tau=100\omega_g$, both normalized to the DC conductivity $\sigma_0$.}

\caption{Comparisons between the Mattis-Bardeen (\full) and our approximate model response function (\dashed) with $\omega_g/2\pi=1$~THz and $\gamma=100\omega_g$. Also shown is the normal state response (\dotted). (a) The time-dependent conductivity. (b) The corresponding time-dependent susceptibility. (c) The frequency-dependent conductivity. Note that the results in (a) and (c) are connected through a Fourier transform.}

\caption{Superconducting-to-normal transmission (a) and reflection (b) ratios for a thin-film superconductor with $\omega_g/2\pi=1$~THz and $\gamma=100\omega_g$ on a dielectric substrate. Results are calculated using the Mattis-Bardeen conductivity (\full), the model frequency-domain conductivity function (\dashed), and by FDTD method using the model time-domain susceptibility function (\chain).}

86 — 1307.5119

\caption[Optional caption for list of figures]{ In \ref{fig:subfig1}, we show the scalar spectral index $n_{s}(k)$, and in \ref{fig:subfig2}, we show the running of the scalar spectral index $\alpha_{s}(k)$, with respect to the momentum scale $k$. The {\bf black} dotted line corresponds to $k_{\Lambda}=0.056~{\rm Mpc}^{-1}$ for $l_{\Lambda}=2500$, the \textcolor{blue}{\bf blue} dotted line corresponds to $k_{L}=4.488\times 10^{-5}~{\rm Mpc}^{-1}$ for $l_{L}=2$, and in all the plots \textcolor{violet}{\bf violet} dashed dotted line represents the pivot scale of momentum at $k_{\star}=0.002~{\rm Mpc}^{-1}$ for $l_{\star}\sim 80$ at which $P_{S}(k_{\star})=2.2\times 10^{-9}$, $n_{S}(k_{\star})=0.96$, $\alpha_{S}(k_{\star})=-0.02$. In both fig~\ref{fig:subfig1} and fig~\ref{fig:subfig2} the PBH formation scale $k_{PBH}$ is lying within the region bounded by the \textcolor{blue}{\bf blue} dotted line and the \textcolor{violet}{\bf violet} dashed dotted line. }

87 — 1307.5282

\caption{\label{D} The coefficient D as a function of the angle $\alpha$ \blue{(in $^\circ$)} in the range of angles considered in Ref. \cite{Waele} \blue{with $\beta=12.5^\circ$}. }

88 — 1307.5653

\caption{\label{tab_caretaker_result}Tracking results of the subway video. The proposed controller improves significantly the tracking performance. The best values are printed in \textcolor{red}{\textbf{red}}.}

\caption{\label{tab_pets_result}Tracking results on the PETS sequence S2.L1, camera view 1, time 12.34. The best values are printed in \textcolor{red}{\textbf{red}}.}

\caption{\label{tab_tud_result}Tracking results for the TUD-Stadtmitte sequence. The best values are printed in \textcolor{red}{\textbf{red}}.}

89 — 1307.6033

\caption{{\color[rgb]{0,0,0}{Symbol Error Rate (SER)}} using $\ell_2/\ell_1$ CS decoder, MMSE decoder and CS-MMSE detector for (i) 2 out of 8 active users and (ii) 4 out of 8 active users.}

\caption{{\color[rgb]{0,0,0}{Symbol Error Rate (SER)}} using $\ell_2/\ell_1$ CS decoder, MMSE decoder and CS-MMSE detector for (i) 6 out of 8 active users and (ii) 8 out of 8 active users.}

90 — 1307.6771

\caption{Illustration of the dynamic simulation box algorithm. When an X-marked projectile approaches the face of the initial simulation box (left panel) by a distance $l\approx \rho_{\max}$ a new simulation box of the same size is generated (right panel) with the particle placed approximately in its geometrical center. % The atoms The \textcolor{red}{position of the} atoms (small shadowed circles) located in the intersection of the old and the new boxes are not changed. In the rest part of the new box the atomic positions are generated anew as described in the text.}

91 — 1307.7285

\caption{Flux driven JPA. (a) Circuit diagram. The transmission line resonator is terminated by a dc SQUID (loop with crosses symbolizing Josephson junctions) at one end. A magnetic flux $\Phi_{\textrm{dc}}+\Phi_{\textrm{rf}}$ penetrating the dc SQUID modulates the resonant frequency. (b) Dependence of the resonant frequency on the dc flux. The red line is a fit of a distributed circuit model~\cite{PhysRevB.74.224506} to the data~(\fullsquare). Blue dot: Operation point used in our experiments. (c) Schematic of the operating principle of the JPA (see text for details).}

\caption{ (a) Signal gain as a function of frequency for different values of the signal power and (b)~signal gain at $f_{\rm pump}/2\,{+}\,10\,\kilo\hertz$ versus signal power. Data: \fullsquare. Line: Guide to the eye. The temperature of the JPA is stabilized at $88\,\milli\kelvin$. }

\caption{ Gain corrected power as a function of the noise source temperature. Red line: Fit to the data (\fullcircle$\!\!$) using~(\ref{eq:BosePlanck}). The JPA temperature is in the range from $92$ to $115\,\milli\kelvin$. }

\caption{(a)~Squeezing level ({\color{blue}\fullsquare}) and photon number ({\color{red} \large\fulltriangle}) plotted as a function of the signal gain when the $30\,\deci\bel$-attenuator is at $50\,\milli\kelvin$. The lines are guides to the eyes. (b)~Photon number as a function of signal power gain in linear units. Data: {\large\color{red}\fulltriangle}. Green dashed line: linear fit. Inset: Signal gain range equivalent to that shown in panel~(a). The two data points with the largest signal gain are excluded from the fit. (c)~Squeezing level ({\color{blue}\fullsquare}) and photon number ({\color{red} \large\fulltriangle}) plotted as a function of the $30\,\deci\bel$-attenuator temperature for $1\,\deci\bel$ signal gain. The lines are guides to the eyes. All error bars are of statistical nature. }

92 — 1307.7382

\caption{\label{mymodels_factor} Factor diagrams for the described models, along with their respective the Gibbs update conditional factorizations for their latent variables. Latent variables are \textcolor{blue}{\textbf{blue}}; the rest are observed. A factor graph is formally defined as a bipartite graph between variables and factors; a factor is visually represented here as a shape surrounding the variables it implicates. These diagrams only show the discrete variables in the models; their Dirichlet-multinomial priors are not shown. }

93 — 1307.7716

\caption{Gain differences between the July and October nights are dominated by temperature variation. Here we see the effect on the average auto correlation before ({\color{orange} orange}) and after (black) application of the temperature dependent gain. \label{fig:auto_compare}}

94 — 1307.7813

\caption{DBG with $k=3$ for the sequences: {\color{red} ACT}{\color{black}GGA}{\color{cyan}GCG} ($awb$) and {\color{red} ACT}{\color{cyan}GCG} ($ab$). The pattern in the sequence generates a $(s,t)$-bubble, from {\color{red}CT}{\color{blue}G} to {\color{cyan}GCG}. In this case, $b=$ {\color{cyan}GCG} and $w=$ GGA have their first letter {\color{blue}G} in common, so the path corresponding to the junction $ab$ has $k-1-1 = 1$ vertex.}

95 — 1308.1876

\caption{\small Maximum eigenvalue of matrix $\tilde{\mathbf{P}}_i^{-1}(l)\mathbf{Q}_i(l)$ against different values of $\lambda_\text{max} \big(\mathbf{X}_i^{-1}(l)\mathbf{Q}_i(l)\big)$. Note that both matrices are rank-1 and positive semidefinite.}\vspace{-0.2cm} \label{fig:lambda} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=\columnwidth]{Rate.eps} \caption{\small The minimum achievable rate amongst all downlink and uplink in different values of SNR.} \label{rate_fig} \end{center} \end{figure} % \begin{figure}[t] % \begin{center} % \includegraphics[scale=0.8]{figures/relative_error.eps} % \caption{\small Relative error of achieved rate (w.r.t minimax bound) against SNR.} % \label{rate_error_fig} % \end{center} % \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=\columnwidth]{iter_Wt.eps} \caption{\small Number of iterations to find $\mathbf{W}$ using Algorithm \ref{Wt_algorithm} in different values of SNR.} \label{iter_Wt_fig} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=\columnwidth]{iterations_LM_BS.eps} \caption{\small Total number of iterations in LM-bisection method versus SNR. Here the number of iterations is calculated by adding up number of iterations in each bisection step.} \label{LM_iter_fig} \end{center} \end{figure} % \begin{figure}[t] % \begin{center} % \includegraphics[scale=0.8]{figures/improvement_percent.eps} % \caption{\small Relative improvement of rate (w.r.t the corresponding rate in case of no BS precoding) against SNR.} % \label{rate_improvement_fig} % \end{center} % \end{figure} \begin{algorithm} \caption{Iterative Algorithm for finding sub-optimal solution of \eqref{MMF-BS1}} \label{Wt_algorithm} \begin{algorithmic} \STATE {\bf initialization:} $\boldsymbol{\Psi}(0)=\sqrt{\frac{P_B}{N_b}}\;\mathbf{I}_{N_b}$, $l \leftarrow 0$, $i \in \mathcal{S}'$ \\ \begin{align*} &A_0=\mathcal{H}\big(\lambda_\text{max}(\tilde{\mathbf{P}}^{-1}_i \mathbf{Q}_i)\big),\;\mathbf{W}(0)=\sqrt{\frac{P_B}{N_b}}~~\mathbf{I}_{N_b} \\ &A_1=\mathcal{H}\big(\lambda_\text{max}(\tilde{\mathbf{P}}^{-1}_i \mathbf{Q}_i)\big),\;\mathbf{W}(0)=\sqrt{\frac{P_B}{\operatorname{Tr}(\boldsymbol{\Lambda}^{-1})}}~~ \mathbf{U}\mathbf{\Lambda}^{-1/2}. \end{align*} \IF {$A_0 \geq A_1$} \STATE $\mathbf{W}(0)=\sqrt{\frac{P_B}{N_b}}~~\mathbf{I}_{N_b}$ % \STATE break \ELSE \STATE $\mathbf{W}(0)=\sqrt{\frac{P_B}{\operatorname{Tr}(\boldsymbol{\Lambda}^{-1})}}~~ \mathbf{U}\mathbf{\Lambda}^{-1/2}$ % \STATE $\mathbf{W}'(0)= \mathbf{H}_0\boldsymbol{\Psi}(0)$ \ENDIF \WHILE{convergence} \STATE $l \leftarrow l+1$ % \STATE find $\boldsymbol{\Psi}(l)$ using \eqref{Wt_power_allocation} \STATE find $\mathbf{W}(l)$ with \eqref{Wt_closed_form} and update $\mathbf{Q}_i(l)$, $\tilde{\mathbf{P}}^{-1}_i(l)$ and $\mathbf{X}_i(l)$ \ENDWHILE \STATE {\bf return $\mathbf{W}$}. \end{algorithmic} \end{algorithm} \begin{table} \begin{center} \caption{Comparison of the complexity of the proposed algorithm with alternating method.} \label{complexity_table} \begin{tabular}{ |c|c|c|c| } \hline % &\multicolumn{2}{ |c| }{Stage} \\ % \hline Method &RS precoding & BS precoding & Total (for $N_r \geq N_b$)\\ \hline \multirow{2}{*}{Alternating \cite{SIDIRO2006}} & SDP-bisection: & SDP-bisection: & $N_a.\,\mathcal{O}\big((N^{2}_r)^6\big)\times$\\ & $\mathcal{O}\big((N^{2}_r)^6\big).\mathcal{O}\big(\sqrt{2N_b}N_r\log (1/\epsilon_{SDP})\big)$ & $\mathcal{O}\big((N^{2}_b)^6\big).\mathcal{O}\big(\sqrt{N_b}N_b \log (1/\epsilon_{SDP})\big)$ & $\mathcal{O}\big(\sqrt{2N_b}N_r\log (1/\epsilon_{SDP})\big)$ \\ \hline \multirow{2}{*}{Proposed} & LM-bisection \cite{DartmannTVT13}: & Algorithm \ref{Wt_algorithm}: & \multirow{2}{*}{} \\ & $N_{LM}.\mathcal{O}((N^{2}_r)^3)$ & $N_w.\mathcal{O}((N^2_r)^3)$ & $(N_w+N_{LM}).\mathcal{O}((N^{2}_r)^3)$ \\ \hline \end{tabular} \end{center} \vspace{-1cm} \end{table} \end{document} }

96 — 1308.3441

\caption{Mode conversion by OLPGs: The intensity distribution of the multi-mode interference (MMI) of a corresponding cw-pump beam equally distributed between the \LPoi- and the \LPoii-mode is shown projected to the xz-plane while the change in modal probe energy is shown on the yz-plane, with z being the propagation direction. The probe pulse intensity distributions at the beginning of the fiber and after maximum conversion are shown in the insets as well as the temporal intensity profiles of both probe modes (red = \LPoi-, blue = \LPoii-mode). The localized parts of the OLPG induced by a femtosecond pump beam are indicated by the grey rectangles. The conversion speed is exaggerated in this schematic representation. See \textcolor{blue}{Media~1} for an animated version of the process.}

\caption{\textbf{(a)} Probe power measured with PD2 as a function of the phase difference between probe and pump beam in front of the fiber at an average pump beam power of 150\,mW and an average probe beam power of 15\,mW. The horizontal dashed line indicates the probe power$P_0$ without a pump beam present. The markers indicate phase delays that are investigated in more detail in sub-figure \textbf{(b):} Here, the pump energy dependent difference between the probe power levels ``with'' and ``without'' pump beam present is depicted for different phase differences measured with PD1 (dotted lines to guide the eye). For details see the text.}

97 — 1308.4080

\caption{Complete NMR characterization of an extract from the laser pyrolysis (combined methanol and DCM extract). The amounts of \protect\includegraphics[scale=0.18]{arom-bond.eps}CH, \protect\includegraphics[scale=0.18]{doublebond.eps}CH$_2$, and \sbond CH$_3$ groups (upper right corner) were determined from the line areas of the $^1$H NMR experiment. The assignment of the different functional groups was carried out by connecting the peaks of the $^{13}$C DEPT135 scan with the peaks of the $^1$H scan via an HSQC experiment.}

\caption{Approximate determination of the relative abundances of \protect\includegraphics[scale=0.18]{doublebond.eps}CH$_2$ and \sbond CH$_3$ groups in the diffuse ISM (see the text for an explanation).}

98 — 1308.5277

\caption{Schematic depicting the initial state method for calculating core-level energy shifts as applied to a PFP monolayer in vacuum, charged, and adsorbed on a Ag(111) three layer surface. The calculated density of states (DOS) for the initial state ({\Large{\textbf{------}}}), charge of the molecule $Q$, external potential $V$ ({\color{red}{{---------}}}), and C1s energy shifts are shown. Occupation of the DOS is denoted by grey regions, with charge added to the molecule in vacuum marked in red. Depictions of the neutral, charged ($Q \approx -0.75$~$e$), and adsorbed PFP on Ag(111) are also shown below, with the change in charge density depicted by isosurfaces of $-0.1$~$e/a_0^3$. C and F atoms are depicted by gray and green balls, respectively. Note the DOS for the PFP monolayer in vacuum was increased by a factor of five for clarity.}

\caption{Charge $Q$ in $e$ of PFP in a pure monolayer ({$\medbullet$}) and a 1:1 mixed PFP+CuPc monolayer ({\color{red}{$\blacksquare$}}) as a function of the number of Ag(111) atomic layers $N$, calculated via a Bader analysis.}

\caption{XPS spectra for multi-layer PFP ({\footnotesize{$\square$}}) and a monolayer of PFP ({\color{red}{{\footnotesize{$\Diamondblack$}}}}) on Ag(111) of the F (left), C$_{\mathrm{F}}$ and C$_{\mathrm{C}}$ (right) atomic species, as shown schematically. C and F atoms are depicted by gray and green balls, respectively. }

\caption{Energy $E$ in eV versus charge $Q$ per molecule in $e$ for (a) C$_{\mathrm{C}}$ and (b) C$_{\mathrm{F}}$ atomic species in PFP and (c) C$_{\mathrm{H}}$ and (d) C$_{\mathrm{N}}$ atomic species in CuPc of the C1s level \(-\varepsilon_\mathrm{C1s}\) in vacuum ($\medcirc$), on an $N$ layer Ag(111) surface ({\color{red}{$\medbullet$}}) and after subtracting the change in external potential due to the Ag(111) surface \(-\varepsilon_\mathrm{C1s} + V\) ({\color{blue}{$\Diamondblack$}}). % provided in Table~\ref{tbl:VextPure}. All energies are taken relative to the binding energy of the neutral molecule $E_{Q=0}$. A quadratic fit to the pure monolayer C1s binding energies in vacuum (\textbf{------}) is also given. }

\caption{Energy $E$ in eV versus charge $Q$ of PFP in $e$ for (a) C$_{\textrm{C}}$ and (b) C$_{\textrm{F}}$ atomic species of the C1s level \(-\varepsilon_\textrm{C1s}\) in a 1:1 mixture with CuPc in vacuum ({\color{red}{$\square$}}), on an $N$ layer Ag(111) surface ({\color{red}{$\blacksquare$}}), and after subtracting the change in external potential \(-\varepsilon_\textrm{C1s} + V\) due to the other molecules in vacuum ({\color{blue}{$\triangle$}}) and due to the $N$ layer Ag(111) surface ({\color{Green}{$\blacktriangle$}}). % provided in Table~\ref{tbl:VextPure}. The binding energies of the pure PFP monolayer adsorbed on an $N$ layer Ag(111) surface ({\color{red}{$\medbullet$}}) are provided for comparison. All energies are taken relative to the binding energy of the neutral molecule $E_{Q=0}$. A quadratic fit to the pure monolayer C1s binding energies in vacuum (\textbf{------}) is also given. The mixed 1:1 CuPc+* structure, where the average external potential ${V} \approx -0.307$~eV is calculated at the positions of the C atoms in PFP, is depicted schematically above. C, N, H, and Cu atoms are depicted by gray, blue, white and orange balls, respectively.}

\caption{The calculated density of states (DOS) (lines) and measured XPS spectra (symbols) for the F1s, N1s, and C1s levels versus binding energy for pure monolayers of PFP ({\color{red}{------}},{\color{red}{$\Diamondblack$}}), CuPc ({\color{blue}{------}},{\color{blue}{$\blacksquare$}}), and a 1:1 mixture of PFP+CuPc (------,$\medbullet$). The PFP and CuPc structures, along with the four different C atomic species, C$_{\textrm{C}}$, C$_{\textrm{F}}$, C$_{\textrm{H}}$, and C$_{\textrm{N}}$, consisting of six and four symmetrically inequivalent C atoms in PFP and CuPc, respectively, are shown above. C, F, N, H, and Cu atoms are depicted by gray, green, blue, white and orange balls, respectively.}

\caption{Comparison of charge $Q$ in $e$ from an effective potential model \(Q^{\textrm{Model}}\) and from DFT calculations \(Q^\textrm{DFT}\) for PFP in vacuum ($\medcirc$), on an $N$ layer Ag(111) surface ({\color{blue}{$\Diamondblack$}}), in a 1:1 mixture with CuPc in vacuum ({\color{red}{$\triangle$}}), and on an $N$ layer Ag(111) surface ({\color{Green}{$\blacktriangle$}}). %Model parameters \(\xi\), \(\zeta\), and \(V\) are provided in Table~\ref{tbl:ModelParameters,tbl:VextPure}. The standard deviations for PFP and PFP+CuPc on Ag(111), $\sigma \approx \pm 0.09$~$e$, are shown as regions of gray. }

\caption{(a) Energy $E$ of the C$_{\textrm{C}}$1s ({\color{red}{$\medbullet$}}) and C$_{\textrm{F}}$1s ({\color{black}{$\blacksquare$}}) levels in eV and (b) charge $Q$ in $e$ ($\medbullet$) of a pure PFP monolayer versus height $h$ in \AA\above a three layer Ag(111) surface. Energies are taken relative to the binding energy of the neutral molecule$E_{Q=0}$ in vacuum. Exponential fits are provided as guides to the eye.}

99 — 1308.5531

\caption{\label{fig:phase-space} \small Representation of the dynamics in a space spanned by the mean spanwise velocity $\langle w \rangle$ and the root-mean-square velocity fluctuations in the wall-normal ($v_{\mathrm{rms}}$) and streamwise ($u_{\mathrm{rms}}$) components for length $L_x=4.7\pi$ and $L_x=4.5\pi$ in (\textit{a}) and (\textit{b}), respectively. The dynamics on the relative attractor within the edge (shown as a full line {\color{black}\solid}) is intermittent in case (\textit{a}) and erratic in case (\textit{b}). In addition, we show the trajectories of the periodic LR state at $L_x=6\pi$ ({\color{red}\solid}) and $L_x=4.6 \pi$ ({\color{red}\dotdashed}), as well as that of the $L$ and $R$ states at $L_x=6\pi$ ({\color{green}\solid}) and $L_x=5\pi$ ({\color{green}\dotdashed}).}

100 — 1308.5702

\caption{\redcol A series of $41\times41$ pixel surface plots demonstrating our PSF fitting of the 2M0838$+$15 BC system. Component A is shown on top, and components BC on the bottom. From left to right: the data, the model, and residuals. \label{fig:psf}}

\caption{ \redcol Top panel: spectral decomposition of the A(BC) components using composite spectral templates from the SpeX prism library as described in section \ref{sect:spt}. The reduced $\chi^2$ values for each spectral fit as a function of A (filled circle) and BC (open triangle) spectral types are shown in the left panel. The composite templates with $\chi^2/\rm{dof} < 1.5$ (grey lines) are shown on the right, in comparison to the unresolved SpeX spectrum of 2M0838+1511 (black line). Contributions from the individual A and BC templates are shown in cyan and pink respectively. The best-fitting A and BC templates are indicated with pink and cyan labels respectively. Note that each point on the left belongs to a unique {\em composite} template, but that individual A and BC template spectra may each contribute to multiple composites. Middle panel: spectral decomposition of the BC components, in a similar fashion to the A(BC) decomposition in the top panel. {\em Bottom panel:} the best fitting spectral templates for the A, B and C components, over plotted on the resolved OSIRIS spectra for components A and BC. \label{fig:spec_a}}

\caption{\redcol An example of the age-mass degeneracies typical for BD systems, with surface gravities and effective temperatures for the 2M0838$+$15 ABC components inferred using empirical trends and evolutionary models, are plotted (from left to right) along a series of isochrones ranging from 300\,Myr to 5\,Gyr. Isochrones (grey lines labeled with$\log{t {\rm [yr]}})$ and lines of constant mass (black lines labeled in units of $M_{\odot}$) using the evolutionary models of \citet{burrows97} are overplotted. Non-singular solutions for the 300\,Myr isochrone are shown with the components labelled by letter. A dynamical mass for the BC system will constrain the system age, temperatures and surface gravities, and allow a direct comparison to model isochrones.\label{fig:iso}}

\caption{Binding energy as a function of total mass for stellar and VLM binary systems. The 2M0838$+$15 A(BC) system is plotted using a 5-point star. A dotted horizontal line makes the minimum binding energy cutoff of $\sim 20\times10^{41}$\,erg observed for the majority of VLM binaries. Most of the other VLM binaries plotted (squares) are taken from the VLM Binary Archive, which is a compilation of data from 144 unique publications which can be accessed at\tt{vlmbinaries.org}. {\rm Additional VLM systems, stellar binaries (crosses) and stellar primaries with VLM secondaries (open circles) are taken from \citet{close90,fischer92,tokovinin97,reid97a,reid97b,reid01,caballero06,daemgen07,kraus07,radigan08,luhman09,faherty10}\label{fig:sf}. Young/low-gravity systems are marked by filled circles and squares.}}

101 — 1309.0453

\caption{\label{fig_fluo}Fluorescence \SI{> 645}{\nm} vs. neutron irradiation time or electron dose. Error bars indicate standard errors. Upper panel: Type 1b HPHT samples. Lower panel: Type IIa CVD samples. Markers: $\bullet\ $Type 1b, neutrons. $\mbox{\tiny$\blacksquare$}\ $Type 1b, electrons, heated during irradiation. Pentagon:~Type 1b, electrons, RT during irradiation. $\circ\ $Type 1b, untreated, for comparison. $+\ $Type IIa CVD, neutrons. $\times\ $Type IIa CVD, electrons, heated during irradiation. $\diamond\ $Sumitomo Type 1b HPHT, neutrons.}

102 — 1309.0493

\caption{This figure, adapted from \cite{Tegetal06}, shows the factors affecting the selection probability $P(\Dell|\Lambda, Q)$ as quantified by the fraction $f(\Lambda, Q, t_0)$ of protons in the form of galaxies at a time $t_0=11$Gyr {\kf a few billion years before} the present age, The displayed curves correspond to $f=.4,.5,.6$. The selection probability $P(\Dell|\Lambda, Q)$ is approximately uniform in the white interior region of the contours shown and decreases quickly outside them. That defines the anthropically allowed range. The most probable values for $\Lambda$ and $Q$ are indicated by the {\color{red}\bigstar}\assuming the no-boundary prior and by the{\color{blue}\bigdot} assuming a uniform one. }

\caption{The $\Lambda$-$Q$ plane. This figure, adapted from \cite{Tegetal06} (the same as Figure \ref{fraction1}) shows the various factors affecting the calculation of the selection probability $P(\Dell|\Lambda, Q)$ that is the basis for anthropic selection. The region that allows for gravitationally collapsed bodies at some time is to the left of the diagonal dotted line assuming the presently observed matter density. The contours trace constant values of the fraction $f(\Lambda, Q, t_0)$ of protons in the form of galaxies at {\pf a time $t_0 \sim11$ Gyr a few billion years before the present age}, as calculated from formulae in \cite{Tegetal06}. The displayed curves correspond to $f=.4,.5,.6$. The selection probability $P(\Dell|\Lambda, Q)$ is approximately uniform in the interior of the contours shown and decreases quickly outside them. The most probable values for $\Lambda$ and $Q$ are indicated by the {\color{red}\bigstar}\assuming the no-boundary prior\eqref{plamq}. These coincide with the observed values. The correlation thus supports the theory that led to it. The most probable values for a uniform prior in $Q$ and $\Lambda$ are at the largest value of $\log \Lambda$ in the allowed range and indicated by the {\color{blue}\bigdot}. These are not close to the observed values.}

103 — 1309.1329

\caption{Vorono\"{i} tessellation of two set of data points, where the \textcolor{red}{red} dots are the seed points: (a) Scattered data set (b) Polygonal mesh after few iterations. }

104 — 1309.1472

\caption{Experimental data. Measured expectation values $\{d^{\rm exp}_j\}$ of the output states in the eigenbasis of the SLD for all the settings $(k,s)$ encompassed in our demonstration. Each row refers to one of the two families of (discordant vs classical) probes, with $s=Q,C$ for rows \abl{a}, \abl{b}, respectively; on the other hand, each column represents a different black box setting $k=1,2,3$ implemented in the protocol. Referring to the ordering of the SLD eigenvectors as given in Table~\ref{tabs2}, the plot legend in all panels is: \textcolor[rgb]{0.2472, 0.24, 0.6}{\ding{108}} $d_1^{\rm exp}$, \textcolor[rgb]{0.6, 0.24, 0.442893}{\ding{110}} $d_2^{\rm exp}$, \textcolor[rgb]{0.6, 0.547014, 0.24}{\ding{117}} $d_3^{\rm exp}$, \textcolor[rgb]{0.24, 0.6, 0.33692}{\ding{115}} $d_4^{\rm exp}$. Error bars amounting to $5\%$ on the measured data as estimated from the pulse imperfections are included.\label{figs5}}

105 — 1309.1724

\caption{Characteristic length $\ell$ vs time $t$, for temperatures: \unit{1080}{\celsius} ({\Large\textcolor{OliveGreen}{$\bullet$}}), \unit{1130}{\celsius} (\textcolor{blue}{$\blacksquare$}), \unit{1180}{\celsius} (\textcolor{red}{$\blacklozenge$}). The dotted lines are linear fits. Inset: semi-log plot of the slopes of the linear fits $b$, as a function of the inverse temperature $1/T$.}

\caption{ Rescaled volumes $v$ of the domains vs their rescaled surfaces $s$, for all isolated domains (\textcolor{cyan}{{\Large$\bullet$}}) and for the percolating domain (\textcolor{OliveGreen}{$\blacklozenge$}) after 16, 32, 45 and \unit{64}{\minute} at \unit{1130}{\celsius}. The error bars indicate the standard deviation of the measures for all boxes of a given size. The black solid line is $v = s^{3/2}$ (ideal spheres), and the black dotted line is $v \sim s$. Some domains shapes are shown as examples.}

\caption{Gaussian curvature distributions $K$, $K_p$, $K_i$ ; before (a), and after rescaling for all domains (b), only percolating (c) and only isolated domains (d) . After 8 (\textcolor{cyan}{{\Large$\bullet$}}), 16 (\textcolor{blue}{$\blacksquare$}), 32 (\textcolor{OliveGreen}{$\blacktriangleleft$}), 45 (\textcolor{magenta}{$\bigstar$}) and \unit{64}{\minute} (\textcolor{red}{$\blacklozenge$}) at \unit{1130}{\celsius}.}

106 — 1309.2321

\caption{\label{fig:AllSingleDiscPlots}{\it Semi-analytically calculated overlap terms for a Single Disc:} The emitter is located at a vertical distance of $15$nm above the disc and moves along $Y_d=0$. The colors correspond to different polarizations of the emitter: $\hat{x}$(\textcolor{red}{{\Large{{{\textbf{\textrm{--}}}}}}}), $\hat{y}$(\textcolor{green}{{\Large{{{\textbf{\textrm{--}}}}}}}) and $\hat{z}$(\textcolor{blue}{{\Large{{{\textbf{\textrm{--}}}}}}})}

\caption{\label{fig:AllDoubleDiscBrightPlots}{\it Semi-analytically calculated overlap terms for the disc dimer bright mode:} The emitter is located on the inversion plane parallel to the discs and moves along $Y_d=0$. The colors correspond to different polarizations of the emitter: $\hat{x}$(\textcolor{red}{{\Large{{{\textbf{\textrm{--}}}}}}}) and $\hat{y}$(\textcolor{green}{{\Large{{{\textbf{\textrm{--}}}}}}}). The $\hat{z}$-polarization has a zero overlap at these resonances.}

\caption{\label{fig:AllDoubleDiscDarkPlots}{\it Semi-analytically calculated overlap terms for the disc dimer dark mode:} The emitter is located on the inversion plane parallel to the discs and moves along $Y_d=0$. The color (\textcolor{blue}{{\Large{{{\textbf{\textrm{--}}}}}}}) corresponds to the $\hat{z}$-polarization of the emitter. The $\hat{x}$ and $\hat{y}$-polarizations produce zero overlap at these resonances.}

107 — 1309.2322

\caption{Premultiplied one-dimensional spectra at $y=0.5h$. \solid, present $Re_\tau=550$; \dotted, present $Re_\tau=180$. {\it (a)} $k_xE_{uu}^{1D}(k_x)$. {\it (b)} $k_zE_{ww}^{1D}(k_z)$; }

\caption{Premultiplied two-dimensional spectra $\phi_{uu}$ of the streamwise velocity, as functions of the streamwise and spanwise wavelengths at three representative wall distances. {\it (a)} Wall units, $y+=15$; {\it (b)} Wall units, $y+=90$ ($y=0.5h$ at $Re_\tau=180$); {\it (c)} Outer units, $y=0.2h$ ($y^+=90$ at $Re_\tau=550$); {\it (d)} Outer units, $y=0.5h$. Shaded contours, $Re_\tau=550$; line contours, $Re_\tau=180$. In all the cases there are five linearly increasing contours. \dashed, locus of two-dimensional isotropic structures $\lambda_z = \lambda_x$; the dotted line in {\it (a)} is $\lambda_x^+ \sim (\lambda_z^+)^3$, passing through $\lambda_x^+ = \lambda_z^+ = 50$; those in {\it (b), (c)} and {\it (d)} are $\lambda_x y = \lambda_z^2$, and the point where this line crosses the dashed one corresponds to three-dimensionally isotropic structures. }

\caption{Premultiplied two-dimensional spectra as functions of the streamwise and spanwise wavelengths at three representative wall distances. {\it (a), (c) (e)}, $\phi_{vv}$; {\it (b), (d) (f)}, $\phi_{ww}$. {\it (a), (b)} Wall units, $y+=15$; {\it (c), (d)} Wall units, $y^+=90$ ($y=0.5h$ at $Re_\tau=180$); {\it (e), (f)} Outer units, $y=0.5h$. Shaded contours, $Re_\tau=550$; line contours, $Re_\tau=180$. In all the cases there are five linearly increasing contours. \dashed, locus of two-dimensional isotropic structures $\lambda_z = \lambda_x$; the dotted lines in {\it (a)} and {\it (b)}, are $\lambda_x^+ \sim (\lambda_z^+)^3$ passing through $\lambda_x^+ = \lambda_z^+ = 50$; those in {\it (c), (d), (e)} and {\it (f)} are $\lambda_x y = \lambda_z^2$ and the point where both lines cross in those figures corresponds to three-dimensionally isotropic structures. }

\caption{Superimposed contours of $0.2$ times the maximum of $\phi_{uu}$ at five wall distances in the outer layer (from dark to light $y=0.1h (0.1h) 0.5h$). They are represented as functions of the streamwise and spanwise wavelengths nondimensionalized with the wall distance. \solid, locus of two-dimensional isotropic structures $\lambda_z/y = \lambda_x/y$; \dashed, $(\lambda_z/y)^2 = \lambda_x/y$. The point where both lines cross corresponds to three-dimensionally isotropic structures.}

\caption{Premultiplied 1--D spectrum of streamwise velocity, in outer units. The shaded contours are the present $Re_\tau = 550$ simulation; the symbols are experiments by \cite{NakNez81}. \solidtrian and \trian, $Re_\tau=696$; \solidsquar and \squar, $Re_\tau=318$. Open symbols, $\bra \lambda\ket_s$; closed symbols, $\bra \lambda\ket_e$. }

\caption{Premultiplied two-dimensional cospectra as functions of the streamwise and spanwise wavelengths at two representative wall distances. {\it (a)} Wall units, $y^+=15$; {\it (b)} Outer units, $y=0.5h$. Shaded contours, $Re_\tau=550$; line contours, $Re_\tau=180$. In all the cases there are five linearly increasing contours. \dashed, locus of two-dimensional isotropic structures $\lambda_z = \lambda_x$; the dotted line in {\it (a)} is $\lambda_x^+ \sim (\lambda_z^+)^3$, passing through $\lambda_x^+ = \lambda_z^+ = 50$; that in {\it (b)} is $\lambda_x y = \lambda_z^2$. }

\caption{Structure parameter $\sigma_{uv}$ as a function of the streamwise and spanwise wavelengths at two representative wall distances. Present $Re_\tau=550$. ({\it a}) wall units, $y^+=15$; ({\it b}) outer units, $y=0.5h$. In both cases there are five linearly increasing contours. \dashed, locus of two-dimensional isotropic structures $\lambda_z = \lambda_x$; the dotted line in {\it (a)} is $(\lambda_z^+)^3 \sim \lambda_x^+$ passing through $y^+=50$; that in {\it (b)} is $\lambda_z^2 = y \lambda_x$. }

108 — 1309.2325

\caption{(\aaa) Near-wall maximum r.m.s. $\overline{u'^2}_{max}^{1/2+}$ of the $u$ fluctuations as a function of $Re_\tau$. \solid, $\overline{u'^{2}}^+_{max} \sim \log(Re_\tau)$. (\bbb) R.m.s. $\overline{u'^2}^{1/2+}$ of the $u$ velocity fluctuations at $y=0.4h$ as a function of $Re_\tau$.\solid, $\overline{u'^{2}}^+ \sim \log^2(Re_\tau)$. \squar, laboratory and atmospheric boundary layers; \trian, laboratory channels; \circle, laboratory pipes; \solidsquar, numerical boundary layer; \solidtrian, numerical channels; \solidcircle, numerical pipes. The size of the big square in (\aaa) represents the uncertainty of the atmospheric measurements it comes from. For boundary layers $h$ is the $95\%$ thickness and for the pipes it is the radius.}

\caption{Two-dimensional spectral densities $\phi^{2D}$ as functions of the wavelengths $\lambda_x/y$ and $\lambda_z/y$. The data come from our numerical turbulent channel at $Re_\tau=950$. The line contours are $\phi^{2D+}_u=0.1$ and each of them comes from a different wall distance \circle, $y^+=100$; \squar, $y^+=200$; \trian, $y^+=300$. The shaded contours are $\phi^{2D+}_v=0.01(\times 3)0.1$, at $y^+=200$. \dashed, $\lambda_z= 2(\lambda_x y/7)^{1/2}$. The $\times$ marks the sizes of the vortex clusters using the equivalence $\Delta_{x,z}=\lambda$ and $\Delta_y = y$.}

\caption{Maximum energy amplifications $G(\lambda_z)$ obtained from the linear stability equations of the mean turbulent profile for fixed disturbance length $\lambda_x = 60h$ and different Reynolds numbers. \dotted, $Re_\tau=200$; \dashed, $Re_\tau=500$; \linesolidtrian, $Re_\tau=10^3$; \chndot, $Re_\tau=2\times10^3$; \solid, $Re_\tau=5\times10^3$; \linesolidsquar, $Re_\tau=10^4$; \dashdash, $Re_\tau = 2\times10^4$. (\aaa) $G(\lambda_z^+)$; the solid vertical line is $\lambda_z^+=100$. (\bbb) $G(\lambda_z/h)$; the solid vertical line is $\lambda_z=3h$.}

109 — 1309.2481

\caption{(a) Comparison of the experimental data of Fig.~\ref{fig:encvelfluxratio} with theoretical predictions. Analytical prediction: horizontal solid line - random foam structure (R) in the narrow channel; dashed line - staircase structure (S.S.), eq.~\eqref{repV_stair}; dotted-dashed line - bamboo structure (B. S.), eq.~\eqref{repV_bamb} ($A =0.19$ cm$^2$, $h=0.1$cm and $a_2=4x/(1+x)$ in cm, with $x=a_2/a_1$). The functions are plotted in the domains of existence of the corresponding structure, as observed experimentally. The vertical dashed lines that separate these domains are not strict frontiers, and some fluctuations are observed. In the ``random domain'', adjacent to the staircase domain, one data point corresponds to a double staircase structure (D.S.), for which a special treatment would lead to a better theoretical prediction. The prediction of eq.~\eqref{eq_vtheo_cor}, determined from the experimental values of the foam structure and the velocity in the wide channel, is given as {\color{red}$\bullet$}. The curved solid line is the velocity ratio expected for a Newtonian fluid \cite{bruus}. The analytical expression is given in Annex. (b) Same data for the fluxes (without the result for a Newtonian fluid). }

110 — 1309.3037

\caption{The intensity independent decay rate, $\alpha$, as a function of applied electric field. A negative electric field points anti-parallel to the \textcolor{Black}{pump Poynting} vector.}

111 — 1309.3598

\caption{Ratios of total execution times of CPU and GPU programs for MC integration and parton-level event generation in $u\bar{d}\to W^++n$gluons with BASES/SPRING~\cite{Kawabata:1995th}. We used an NVidia Tesla C2075 GPU with CUDA\trademark 4.2 and an Intel\registered Core i7 2.67 GHz. \label{tab:gpu}}

\caption{Summary of computing requirements for NNLO calculations. Numbers were obtained on Intel\registered Xeon\registered CPU's with varying clock frequency and are therefore not directly comparable. \label{tab:nnlo_requirements}}

112 — 1309.3990

\caption{\label{fig:powerspec} \footnotesize{\bfseries Two dimensional Brownian motion and its variation with trap laser intensity.} All displayed data are for 105.1~nm spheres at 1~mbar. {\bfseries a,} Autocorrelation function (ACF) of experimentally measured radial position (blue points with grey error bars) and fit with theory (black solid line) (\blue{supplementary information equation S4}). {\bfseries b,} Measured axial (black points) and radial position (blue points) power spectra with fits to $P(\omega)$ (solid lines) over trap frequency ${\omega \over 2 \pi}$. Because of the low damping at mbar pressures, the axial ($z$) and radial ($x$) peak frequencies, $\tilde{\omega}_{x, z} = \sqrt{\omega_{x, z}^2 - {\Gamma_{\sf CM}^2/2}}$, approximate the trap frequencies $\omega_{x, z}$. {\bfseries c,} Measured radial power spectra (points) with fits (solid lines) for three values of trapping laser intensity. From left to right the intensity of the trapping laser beams is increased by a factor of 2.1 and the peak frequency ${\tilde{\omega}_x/2 \pi}$ increases with intensity, as expected. If there were no temperature increase from the initial intensity (red) the peak height would behave as indicated by the grey solid lines. However, it can be seen that the height of the peak, $A/\Gamma_{\sf CM} \omega_x^2 = 2 k_B T_{\sf CM}/ M \Gamma_{\sf CM} \omega_x^2$, increases with intensity in comparison to the grey lines, indicating the increase of the centre-of-mass temperature, $T_{\sf CM}$, of the sphere.}

113 — 1309.4180

\caption{The SEDs (left), light curves (middle) and electron distributions (right) in the pure Synchrotron Self-Compton (SSC) scenario with change of the magnetic field as the cause of the flare. {\bf Left:} The colored triangles are simultaneous data points from SMARTS, \swift and \fermi, with red being the high states and blue being the low states. The open triangles are from the flare 1 identified by \cite{chatterjee_2013:0208_opticalonly}, while the filled triangles are from flare 2. The open blue triangles are from MJD 54720-54730; the open red triangles are from MJD 54730-54740; the filled blue triangles are from MJD 55160-55170; the filled red triangles and bow-tie (SMARTS and \swift), as well as the red \fermi upper-limits are from MJD 55190-55200. The grey squares include \swift, \planck and ground based radio data in November of 2009 reported by \cite{giommi_2012:planck}. The gray circles are \wise data taken in MJD 55367-55369 (around June 23, 2010). The three histograms show the SEDs before, during, and after the peak of a flare, with legends in corresponding colors showing the simulation time in the observer's frame. The dotted magenta line is the median thermal emission from radio loud quasars \cite{elvis_etal94_qsr_sed} scaled according to the observed UVOT flux of this source. It is added to the simulated SEDs in the post-processing as a steady component. The dashed red line is the isolated second order SSC emission during the peak of the flare, while the dashed grey line shows the first order SSC. {\bf Middle:} In the bottom panel the open circles show the 10-day-averaged optical light curves in B (blue) and J (magenta) bands starting from MJD 55110. The data points are connected by dotted lines to guide the eyes. The histograms show two simulated synchrotron light curves at similar frequencies. In the upper panel three simulated IC light curves are shown, with green, orange, and black solid lines representing the energy bands in \xray, \fermi \gray, and very high energy (VHE) \gray. The shaded gray areas mark the phase when the simulation is still in setup phase. The vertical dotted line marks the peak of the synchrotron light curves. {\bf Right:} The electron distributions in the front-center zone. The simulation time shown are based in the frame of the emitting blob. }

\caption{The SEDs (left), light curves (middle) and electron distributions (right) of the EC/dusty torus scenario with change of the magnetic field as the cause of the flare. The data points are similar to those in \ref{fig:sscb}, except that in the SED the blue dashed line shows the EC component, and in the light curves there is no VHE \gray because of zero flux.}

114 — 1309.4385

\caption{\textbf{Movie} Frames from a depth-map movie (\textcolor{blue}{Media 1}), of a three-dimensional pendulum consisting of a baseball suspended by a $170$ cm rope swinging through a $25$ degree solid angle. The transverse resolution is $32\times 32$ pixels with a frame rate of $14$ frames per second.}

115 — 1309.4724

\caption{\label{fig:concept} (\red{C}olor online) Conceptual scheme of a heralding qubit amplifier. Input state is transformed according to Eq. (\ref{eq:amplification}). D -- detector, EPR -- ancillary photons, G -- amplifier, FF -- feed forward.}

\caption{\label{fig:scheme} (\red{C}olor online) Scheme for state-\red{dependent} linear-optical qubit amplifier as described in the text. EPR -- source of entangled ancillary photon pairs, PBS -- polarizing beam splitter, PPBS -- partially polarizing beam splitter (defined in the text), WP -- wave plate, PDF -- polarization dependent filter, D -- standard polarization analysis detection block (for reference see \cite{halenkova2012detector}).}

\caption{\label{fig:psucc_fid_Infgain} (\red{C}olor online) Success probability \red{$P_\mathrm{succ}$ given by Eq.~(\ref{eq:r0Psucc})} as a function of output state fidelity \red{$F_\mathrm{QFF}$ given by Eq.~(\ref{eq:r0fidelity})} in the case of infinite gain is depicted for four different input states as described in the text.}

\caption{\label{fig:psucc_fid_gain} (\red{C}olor online) Success probability \red{$P_{\mathrm{succ}}$} as a function of both output state fidelity and amplification gain \red{$G_{\mathrm{FF}}$} is depicted for four different input states as described in the text. THR stands for threshold of unreachable area.}

\caption{\label{fig:spheres} (\red{C}olor online) Probability density function $g = g(\theta,\kappa)$ \red{given by Eq.~(\ref{eq:density})} over the Poincaré sphere for various values of $\kappa$ parameter used in subsequent numerical simulations. Labels $|H\rangle$, $|D\rangle$ and \red{$|R\rangle=(|H\rangle +i|V\rangle)/2$} denote position of horizontal, diagonal and right-hand circular polarization states respectively.}

\caption{\label{tab:kent}Values of medians and first deciles of the \red{the Von Mises-Fisher} distribution for several values of \red{the concentration} parameter $\kappa$.}

\caption{\label{fig:psucc_fid_kappa} (\red{C}olor online) \red{Maximum achievable success probability $\langle P\rangle$} as a function of average fidelity \red{$\langle F\rangle$} for various values of average overall gain $\langle G\rangle$ and state knowledge described by parameter $\kappa$ of probability density function $g = g(\theta,\kappa)$ \red{given by Eq.~(\ref{eq:density})}.}

\caption{\label{fig:kappa_merit} (\red{C}olor online) Merit function $M$ \red{given by Eq.~(\ref{eq:merit})} depicted for various parameters $\kappa$ and average gains $\langle G\rangle$.}

116 — 1309.4890

\caption{DYON simulation results for a pure deuterium plasma or with \textcolor{correction}{wall-sputtering} models i.e. carbon wall or beryllium wall. Each line indicates the electron power losses due to the radiation and ionization: carbon wall (dotted blue), beryllium wall (dashed red), and pure deuterium plasma (solid black). In the case of the carbon wall, the first peak (mainly from deuterium radiation) is much smaller than the second peak, which results from carbon impurities.}

\caption{\textcolor{correction}{Wall-sputtering and recycling} models used in DYON simulations with the ITER-like wall }

\caption{\textcolor{correction}{These figures compare the photomultiplier tube data ($Be^{1+}$ ($527 [nm]$)) between the measured data and the synthetic data, and show the significance of the new models used in the DYON simulations. The red solid lines are the photomultiplier tube data in JET for $\# 82003$. The blue solid line in (a) is the synthetic data with the condition given in Table \ref{conditionforILWsim} (i.e. with physical sputtering model on BeO layer, erosion model of BeO layer, and initial Be content). The black dashed line and solid line in (b) are without initial Be content or any physical sputtering model, respectively. The black solid line in (c) is for pure Be wall. The black solid line in (d) is without erosion model of BeO layer i.e. continuous BeO layer.} }

\caption{\textcolor{correction}{The figures show the effects of the deuterium recycling coefficients. (a) electron density (b) number of photons emitted from $D^{0}$(D alpha) ($465[nm]$). The red lines are (a) Thomson scattering data (b) the measured photomultiplier tube data in JET ($\# 82003$). The blue and black lines are the DYON simulation results with the growing model or decay model of $Y^D_D$, respectively.} }

\caption{\textcolor{correction}{The figures show the effects of fuelling efficiency. (a) D atom puffing with the assumed fuelling efficiency (black $30\%$, blue $10\%$, and green $0\%$), (b) electron densities, obtained by DYON simulation (black $30\%$, blue $10\%$, and green $0\%$) and measured by Thomson scattering (red) in JET ($\# 82003$).}}

\caption{\textcolor{correction}{The figures show the effects of $n^0_c(0)$ on the radiation barrier and the consitituent radiated power. (a) bolometry data (red) in JET ($\# 82003$), and the simulated radiated power in the DYON simulations (solid black $1 \%$, solid blue $0.5 \%$, and solid green $0 \%$ of $n^0_c(0)$). (b) the constituent radiated power (solid blue: total radiated power, dashed red: Be, dashed green: D, dashed black: C, dashed blue: O). For the simulation in (b), the $n^0_c(0)$ is assumed to be $0.5\%$, as given for the simulation in Figure \ref{Demo82003a} and \ref{Demo82003b}. } }

\caption{The cyan lines show the Townsend criterion at different effective connection lengths, as indicated with $500$ and $2000 [m]$, respectively. The black, red, and blue lines indicate the criterion for plasma burn-through, i.e. the minimum electric field for plasma burn-through in the case of a pure deuterium plasma (black), beryllium wall (red), and carbon wall (blue), respectively. The \textcolor{correction}{wall-sputtering} models described in section \ref{ReviewofPSImodelsforcarbonwall} and \ref{NewPSImodelsforITER-Likewall} are used for the simulations, and the required plasma parameters are given by Table \ref{defaultvalues}. The area above both the burn-through criterion and Townsend criterion represents the operation space available for successful start-up in JET. \cor{The red circles indicate the successful plasma burn-through in JET experiments with ITER-like wall ($\# 80239 \sim 82905$).} }

117 — 1309.4907

\caption{\color{Blue} Comparison of the performance of the observer under different constant number of iterations $q\in \{20,50,100,300\}$ on one side and under the proposed updating scheme that start at $q(0)=20$. {\bf (a)} Case where the observer knows the exact value of the parameter $a=10$. {\bf (b)} Case where the observer uses an erroneous value $\hat a=7$ instead of $a=10$. The radius of each circle is equal to twice the corresponding variance.}

118 — 1309.5258

\caption{\label{cont_dual} Continuum extrapolation of the trace anomaly at $T\approx 214$~MeV with (blue points) and without (\pred points) tree level improvement. The left panel shows our results with $2+1+1$ flavors of 4-step stout improved staggered quarks extrapolating from $N_t=6,8,10$ and $12$, whereas on the right panel $N_f=2+1$ 2-step stout improved quarks are used on $N_t=8,10,12$ and $16$. }

\caption{\label{normalization}{\em Left:} contributions of the light (magenta) and strange quarks (turquoise) to the the pressure at $T=214$~MeV at our two finest lattice spacings. The curves represent a scan though various theories with different masses. The sum of the area under the curves gives $p/T^4$. {\em Right:} continuum extrapolation of the pressure at $T\approx 214$~MeV with (blue) and without (\pred) tree level improvement. Only statistical errors are shown. }

119 — 1309.6623

\caption{\label{fig:8x8zerosD40} Zeros of Real (\textcolor{blue}{$\blacksquare$}) and Imaginary (\textcolor{red}{$\Box$}) part of the partition function of Ising model at the volume $8\times 8$ from the HOTRG calculation with $D_s=40$ are on the exact solution lines. Gray dots: MC reweighting solution. Thick Black curve: the region of confidence for the MC reweighting result, above this line, the MC error is large.}

120 — 1309.6901

\caption{Recorded transverse voltage on the GGG/YIG($\SI{61}{\nano m}$)/Pt($\SI{11}{\nano m}$) sample as a function of the external magnetic field strength for $\alpha=45^\circ$. \red{The arrows indicate the sweep direction of the external magnetic field in the experiment.} \textbf{(a)} For $I_\mathrm{d}<0$ a positive offset voltage signal is recorded that exhibits the typical features of the spin Hall magnetoresistance. \textbf{(b)} Reversing the direction of $I_\mathrm{d}$ also inverts the observed voltage signal. \textbf{(c)} Adding $V_\mathrm{t}(+I_\mathrm{d})$ and $V_\mathrm{t}(-I_\mathrm{d})$ %and dividing the result by two reveals the much smaller, thermal (spin Seebeck) component. The large spikes close to the YIG's coercive fields are attributed to either domain reconfiguration or spin torque effects. The inset shows $V_\mathrm{iSSE}$ at large fields for $\alpha=0^\circ$. Here $V_\mathrm{iSSE}$ stays constant for fields of up to $\SI{7}{T}$. }

\caption{\textbf{(a)} Thermal component ($V_\mathrm{iSSE}$) of the recorded voltage (full blue symbols) and Pt temperature (open circles) as a function of the square of applied current on the GGG/YIG($\SI{61}{\nano m}$)/Pt($\SI{11}{\nano m}$) sample. To obtain the individual $V_\mathrm{iSSE}(I_\mathrm{d})$ data points the external magnetic field was rotated at a fixed field strength of $\SI{1}{T}$, from which the spin Seebeck voltage is extracted [$V_\mathrm{iSSE}\equiv V_\mathrm{iSSE}(\alpha=0^\circ)$] for each value of the driving current $\pm I_\mathrm{d}$. The observed $V_\mathrm{iSSE}$ scales quadratically with $I_\mathrm{d}$ as does $T_\mathrm{Pt}$. \red{The indicated error for $T_\mathrm{Pt}$ is an upper estimate for the uncertainty in the fitting algorithm and accounts for the fact that the two rightmost data points were extrapolated. The inset shows the resistivity of the Pt film as a function of temperature.} \textbf{(b)} Generated spin current per applied heating power for the samples investigated in this letter. The numbers in parentheses give the layer thicknesses in $\SI{}{\nano m}$. Taking the smaller spin mixing conductance of the YIG/Au interface into account, all samples give very similar spin current generation efficiencies.}

121 — 1309.6990

\caption{(a) Main: Dispersion relations when the forcing amplitudes is increased (from top to bottom): $p$ goes from 3 (top) to 100 (bottom) Pa. $P_s=0$. Dashed lines correspond to $\omega^2=\frac{D}{\rho}k^5+\frac{T}{\rho}k^3$ with $T$ the only fitting parameter: $T=4$ to 40 N/m (from top to bottom). Inset: same figure with $P_s=130$ Pa ($T_s=22$ N/m), with $10 \leq p \leq 130$ Pa. Dashed line, best fit: $T=22$ N/m (b) Main: $T/T_s$ as a function of $p/P_s$, for various $P_s=$ ({\color{red}$\square$}) $P_s^0$, ({\color{blue} $\square$}) 7, ({\color{green}$\square$}, {\color{green}$\circ$}) 55, ({\color{magenta}$\square$}, {\color{magenta}$\circ$}) 130, ({\color{cyan}$\square$}, {\color{cyan}$\circ$}) 274 Pa for $R=10$ cm, and $P_s=P_s^0=4$ ({\color{red}$\star$}) Pa for the $R=20$ cm. These values correspond, from Eq.\(\ref{stth}), to $T_s=$4, 9, 17, 22, 39 N/m, and $T_s=5$ N/m, respectively. Dashed lines correspond to $T/T_s=1$ for $p/P_s<1$, and $T/T_s\sim (p/P_s)^{0.4}$ for $p/P_s>1$. Inset: same figure in dimensional units. For all colors, $(\square,\star)$ indicate depression and $(\circ)$ overpressure.}

122 — 1309.6992

\caption{(Color online) Main: Capillary exponent $\alpha$ as a function of $\epsilon_I$ for various fluids $\nu=1.1\ 10^{-7}$ ($\star$), $10^{-6}$ ({\color{red}$\diamond$}), $2\ 10^{-6}$ ({\color{red}$\square$}), $3\ 10^{-6}$ ({\color{blue}$\square$}), $4\ 10^{-6}$ ({\color{green}$\square$}), $5\ 10^{-6}$ (GW) ({\color{black}$\square$}), $5\ 10^{-6}$ (oil) ({\color{black}$\circ$}), and $10^{-5}$ m$^2$/s ({\color{magenta}$\circ$}). Inset: $\alpha$ vs. $\nu$ for fixed forcing $\epsilon_I\approx 5\ 10^{-5}$ m$^3$s$^{-3}$. The theoretical capillary exponent $\alpha=17/6$ is indicated in dashed (red) lines.\label{spbilan2}}

123 — 1309.7308

\caption{\label{fig:rietveld}(color online) %Main panel: {\redtwo Best-fit Mn-O-Mn bond angles (bullets) and Mn-O bond lengths (triangles) %from the Rietveld refinement in \LSMO\as a function of$x$ (left panel) and in \oLCSMO\as a function of$y$ (right panel). } The dashed lines are guides to the eye. In the Pnma structures ($y>0.6$), angles and lengths of inequivalent bonds are marked by open symbols, while filled symbols mark average quantities. }

124 — 1309.7791

\caption{\it DNA-chart of CMFV models. Yellow means \colorbox{yellow}{enhancement}, black means \colorbox{black}{\textcolor{white}{\bf suppression}} and white means \protect\framebox{no change}. Blue arrows \textcolor{blue}{$\Leftrightarrow$} indicate correlation and green arrows \textcolor{green}{$\Leftrightarrow$} indicate anti-correlation. }

\caption{\it DNA-chart of $U(2)^3$ models. Yellow means \colorbox{yellow}{enhancement}, black means \colorbox{black}{\textcolor{white}{\bf suppression}} and white means \protect\framebox{no change}. Blue arrows \textcolor{blue}{$\Leftrightarrow$} indicate correlation and green arrows \textcolor{green}{$\Leftrightarrow$} indicate anti-correlation. }

\caption{\it DNA-charts of $Z^\prime$ models with LH and RH currents. Yellow means \colorbox{yellow}{enhancement}, black means \colorbox{black}{\textcolor{white}{\bf suppression}} and white means \protect\framebox{no change}. Blue arrows \textcolor{blue}{$\Leftrightarrow$} indicate correlation and green arrows \textcolor{green}{$\Leftrightarrow$} indicate anti-correlation. }

125 — 1310.0118

\caption{Cosmological constraints on the model parameters, $m_a$ and $f_a$ with $T_R=10^{-2}\GeV, 1\GeV, 10^3\GeV$ and $10^6\GeV$. \red{Here we use the BBN constraint \eqref{bh-3} for $B_h=10^{-3}$}. Colored region is excluded by BBN or gravitino overproduction. \red{The abundance of the gravitino produced from thermal plasma(R-axion decay) exceeds that of DM in the blue(red) region. Here we have set $m_{\tilde g}=10 {\rm TeV}$. }}

\caption{\red{Cosmological constraints on the model parameters, $m_a$ and $f_a$ with $T_R=10^{-2}\GeV, 1\GeV, 10^3\GeV$ and $10^6\GeV$. Here we use the BBN constraint \eqref{bh1} for $B_h=1$. Colored region is excluded by BBN or gravitino overproduction. The abundance of the gravitino produced from thermal plasma(R-axion decay) exceeds that of DM in the blue(red) region. Here we have set $m_{\tilde g}=10 {\rm TeV}$. }}

126 — 1310.0220

\caption{Growing DSM2 interfaces and one-point distribution. (a,b) Snapshots of a circular (a) and flat (b) interface. Elapsed time after emission of laser pulses is indicated. (c) Skewness $\expct{h^3}_{\rm c} / \expct{h^2}_{\rm c}^{3/2}$ and kurtosis $\expct{h^4}_{\rm c} / \expct{h^2}_{\rm c}^2$ (symbols), compared with the values for the GUE and GOE Tracy-Widom distributions (dashed and dotted lines). (d) One-point distribution of the rescaled local height $q \equiv (h - v_\infty t)/(\Gamma t)^{1/3}$ (symbols), compared with the GUE and GOE Tracy-Widom distributions (dashed and dotted lines). The data for the circular interfaces (solid symbols) were measured at $t = 10\unit{s}$ (blue \fullcircle) and $30\unit{s}$ (red \fulldiamond), while those for the flat interfaces were taken at $t = 20\unit{s}$ (turquoise \opentriangle) and $60\unit{s}$ (purple \opentriangledown). Note that the standard random variable for the GOE Tracy-Widom distribution is multiplied by $2^{-2/3}$ here, as suggested by analytical studies. All panels are reprinted from \cite{Takeuchi.Sano-JSP2012} with adaptations, with kind permission from Springer Science+Business Media.}

127 — 1310.0509

\caption{Three COD matrices correspond to the {\color{ddblue}three red dotted paths} on the trees above}

128 — 1310.1314

\caption{An example for linear deterministic IRC with $n_d = 2$, $n_c = 3$, $n_r=6$, and $n_s=5$. The scheme is shown for time slot $b$. Only RX$_2$ is shown \textcolor{c5}{for} clarity.}

129 — 1310.2385

\caption{\textcolor{C1}{Each} base station \textcolor{C1}{serves} the users located in its cell. However, its signal can be received over the noise level in some areas of the adjacent cells. For example, the users in cell 2 and 3 receive an interference signal from the base stations in cell 1 and 2, respectively. On the right, the topology of this network is shown. \textcolor{C1}{Note that each receiver experiences at most one interference.}}

130 — 1310.2521

\caption{\label{fig:fig2} Extracting the boson coupling mode dispersion assuming the following scattering $\boldsymbol{q}$ directionalities in (a): horizontal and vertical (H/V), intra-hole-pocket (Intra), and inter-hole-pocket (Inter). (b) $|\Omega_{\text{kink}}|$ and $\Omega_{\text{boson}}^{*}$ as a function of FS angle $\theta$. $\Omega_{\text{boson}}^{*}$ {[}Equation (\ref{eq:eq3}){]} is the boson energy for the special cases of scattering along symmetry directions such that $\Delta(\boldsymbol{k'})=\Delta(\boldsymbol{k})$, as depicted in (a). (c), (d) Dispersions of $|\Omega_{\text{kink}}(\boldsymbol{q})|$ (red \fullsquare{}) and $\Omega_{\text{boson}}^{*}(\boldsymbol{q})$ (blue \fullcircle{}), assuming that scattering occurs horizontally/vertically in the Brillouin zone. Curves for $|\Omega_{\text{kink}}(\boldsymbol{q})|$ and $\Omega_{\text{boson}}^{*}(\boldsymbol{q})$ have been extracted considering the short (c) and long (d) H/V scattering channels separately. The extracted dispersions are compared to those of Cu-O phonons measured by INS in YBa\textsubscript{2}Cu\textsubscript{3}O\textsubscript{7} (\opentriangle{}) and YBa\textsubscript{2}Cu\textsubscript{3}O\textsubscript{6} (\opentriangledown{}) \cite{Reichardt1996}, as well as phonons observed by IXS in Bi2201 (\opendiamond{}) \cite{Graf2008}. The pink highlighted Cu-O bond-stretching (Cu-O BS) branch, in particular, has been identified by some previous experiments as possibly relevant to the nodal kink. (e), (f) Analogous plots assuming diagonal intra- and inter-pocket scattering. The green shaded line in (f) is the approximate dispersion of the high-energy branch of spin fluctuations (SF) observed in optimally-doped Bi2212 \cite{Xu2009}. The hatched area indicates that this region is observed in the neutron data to be somewhat filled in by the width of the dispersion peaks. For simplicity, error bars from Figures \ref{fig:fig1}(e) and (f) are shown only for the $\Omega_{\text{boson}}^{*}(\boldsymbol{q})$ curve in each panel.}

131 — 1310.2968

\caption{Stratigraphic fossil sampling across three different fossiliferous horizons. (A) An example of the four different sets of fossil samples. Three sets of fossils were assembled by selecting all of the available fossilization events contained within each of the three intervals: $\mathcal{S}_1$, $\mathcal{S}_2$, and $\mathcal{S}_3$. The fourth set of fossils was constructed by sampling 50\% of the fossils in horizon $\mathcal{S}_1$ and half of the fossils in $\mathcal{S}_3$. (B) The coverage probability as a function of the true age for estimates using each of the four sets of fossils: $\mathcal{S}_1$ ($\newmoon$), $\mathcal{S}_2$ ($\times$), $\mathcal{S}_3$ ($\square$), and $\mathcal{S}_1+\mathcal{S}_3$ ({\textcolor{purpcol}{$\boldsymbol{\bigotimes}$}}). (C) The credible interval width for increasing node ages.}

\caption{The effect of non-random extant species sampling on the coverage probability and precision of node age estimates under the FBD model with 10\% of the fossils sampled at random. (A) The coverage probabilities and (B) credible interval widths compared to the true node age when extant species are sampled to maximize the diversity (deep nodes; $\newmoon$), for samples emulating outgroup sampling with significantly imbalanced root nodes ($\times$), and when all extant species are sampled ({\textcolor{red}{$\boldsymbol{\bigotimes}$}}).}

132 — 1310.3283

\caption{\label{fig:3struct_fE} \emph{Electron spectrum for different targets from 3D simulations. There are 3 types of targets: (}\textcolor{blue}{\emph{a}}\emph{) a flat target with $1\mu m$ pre-plasma; (}\textcolor{blue}{\emph{b}}\emph{) a target with slab structures on the front, the depth and width of the slabs are $10\mu m$ and $1\mu m$. The spacing between the slabs is $2\mu m$; (}\textcolor{blue}{\emph{c}}\emph{) a target with tower structures on the front, the depth and width of the towers are also $10\mu m$ and $1\mu m$. The spacing between towers in both transverse directions is $2\mu m$. (}\textcolor{blue}{\emph{d}}\emph{) shows the electron energy spectra of targets (}\textcolor{blue}{\emph{a}}\emph{), (}\textcolor{blue}{\emph{b}}\emph{) and (}\textcolor{blue}{\emph{c}}\emph{). The blue curve for target (}\textcolor{blue}{\emph{a}}\emph{) is commonly observed in experiments on flat targets. The green and red curves for targets (}\textcolor{blue}{\emph{b}}\emph{) and (}\textcolor{blue}{\emph{c}}\emph{) show substantial increases in the yield of electrons above $50 MeV$.}}

\caption{\label{fig:ang_distrib_3D} \emph{Angular distribution of fast electrons ( $>1MeV$). The top three graphs are fast electron number distributions (on a color log scale) as a function of kinetic energy and angle $\theta$ with respect to the forward (z) direction. The bottom three graphs are fast electron number distributions as a function of direction in the xz and yz planes. (}\textcolor{blue}{\emph{a1}}\emph{), (}\textcolor{blue}{\emph{a2}}\emph{) correspond to the flat target in Fig. }\textcolor{blue}{\emph{\ref{fig:3struct_fE}(a)}}\emph{. An overall divergence angle of about $60^{\circ}$ is seen. (}\textcolor{blue}{\emph{b1}}\emph{), (}\textcolor{blue}{\emph{b2}}\emph{) correspond to the slab-structured target in Fig. }\textcolor{blue}{\emph{\ref{fig:3struct_fE}(b)}}\emph{. The over all divergence angle is about $40^{\circ}$. (}\textcolor{blue}{\emph{c1}}\emph{), (}\textcolor{blue}{\emph{c2}}\emph{) correspond to the tower-structured target in Fig. }\textcolor{blue}{\emph{\ref{fig:3struct_fE}(c)}}\emph{. The distribution shows two peaks at $\theta_{y}\approx\pm4-5^{\circ}$. For electrons $>100MeV$, each cone angle is about $4-5^{\circ}$. }}

\caption{\label{fig:physics_2D} \emph{Electron injection and acceleration mechanisms as revealed in a 2D simulation. (}\textcolor{blue}{\emph{a}}\emph{) is a plot of electron density on log scale when the peak of the laser pulse is located at $Z=-4\mu m$. Electron bunches are pulled out from the structure surfaces by the laser E field. (}\textcolor{blue}{\emph{b}}\emph{) is a cartoon showing the accelerating and guiding mechanisms of the electrons pulled out into the gaps. The red ellipse represents the laser pulse. The black dots are the electrons. The trajectories shown here are substantially different from those revealed using fully 3D simulations (see Fig. \textcolor{blue}{\ref{fig:track_GRT_SPK}}).}}

\caption{\label{fig:track_GRT_SPK} \emph{Trajectories of the higher-energy hot electrons ($>120MeV$) from 3D simulations. (}\textcolor{blue}{\emph{a1}}\emph{), (}\textcolor{blue}{\emph{a2}}\emph{), (}\textcolor{blue}{\emph{a3}}\emph{) correspond to the slab target Fig. }\textcolor{blue}{\emph{\ref{fig:3struct_fE}(b)}}\emph{. (}\textcolor{blue}{\emph{a2}}\emph{), (}\textcolor{blue}{\emph{a3}}\emph{) are side-views while looking along either y-axis or x-axis. Red curves indicate electrons with initial positions $X>0$, while blue are those from $X<0$. In the case of slabs, electrons bounce back and forth between the structure surfaces in the x direction, and are bent towards the center in the y direction. (}\textcolor{blue}{\emph{b1}}\emph{), (}\textcolor{blue}{\emph{b2}}\emph{), (}\textcolor{blue}{\emph{b3}}\emph{) correspond to the tower target Fig. }\textcolor{blue}{\emph{\ref{fig:3struct_fE}(c)}}\emph{. Electrons are guided by the structure surfaces in the x direction, and pinch in the y direction. These trajectories are considerably different from those predicted using 2D simulations (see Fig. \textcolor{blue}{\ref{fig:physics_2D}})}}

\caption{\label{fig:physics_fin} \emph{Guiding mechanisms in 3D slabs. In (}\textcolor{blue}{\emph{a}}\emph{), in the x-direction, electrons are bounced back and forth between surfaces, while in the y-direction, they are guided towards the center. (}\textcolor{blue}{\emph{b}}\emph{) is a plot of the B field perpendicular to the laser direction at $Z=-5\mu m$, where the peak of the laser pulse is located. A spatial average over $\lambda$ in z direction is applied to minimize the oscillating $B_{y}$ field from the laser. $B_{\perp}$ is normalized by the peak $B$ field of the laser $B_{0}$. The color scheme indicates the magnitude while the arrows point out the directions. The lengths of the arrows are proportional to the magnitude of $B_{\perp}$. }}

\caption{\label{fig:physics_spk} \emph{Guiding mechanisms in 3D towers. In (}\textcolor{blue}{\emph{a}}\emph{), The quasi-static B field induced by the hot electron current and the return current is indicated by the green arrows. Since the laser polarization is in the x direction, the electrons pulled out by the laser E field mainly fill in the gaps in the XZ plane rather than the YZ plane. The $B_{x}$ field that dominates in the gaps pulls the electrons along y towards the center. In the x direction near the surfaces, the $E_{x}$ and $B_{y}$ fields still balance out. So electrons are constrained in the x direction and pinched in the y direction. Similar to Fig. }\textcolor{blue}{\emph{\ref{fig:physics_fin}(b)}}\emph{, (}\textcolor{blue}{\emph{b}}\emph{) is a plot of $B_{\perp}$ at $Z=-5\mu m$ where the peak of the laser pulse is located.}}

133 — 1310.3662

\caption{Sensing matrix in units of [V/\microrad] (pitch). All elements measured at 9.7\,Hz in the closed-loop system, but with the feedback at 9.7\,Hz notched out. Numbers in\textcolor{gray}{gray} are the measurement results that have coherence less than 0.9. Boxes highlight the elements actually used for computing the control servo's input matrix (inverse of sensing matrix); all other elements are set to zero.}

134 — 1310.4455

\caption{Correlation plot of $|(M_\nu)_{ee}|$ vs.\$|(M_\nu)_{\tau\tau}|$ based on the global fit results of Forero \textit{et al.}~\cite{Forero} assuming a normal neutrino mass spectrum and allowing $m_0$ to vary between zero and $0.3\,\text{eV}$. The boundaries of the allowed areas are depicted by the following symbols: best fit: {\large $\ast$}, $1\sigma:$ \textcolor{red}{$\blacktriangle$}, $3\sigma:$ \textcolor{blue}{$\bullet$}.}

135 — 1310.4835

\caption{Upper: The logarithmic EWs' distribution as a function of \lya\fluxes for all candidates (empty black squares), targeted candidates (empty green and blue squares are targets in NB665 and NB673 images, respectively), and the spectroscopically confirmed LAEs (red filled squares) at z=4.5. Lower: The histogram distribution of log EWs for all candidates (black empty histogram), targets (cyan line-filled histogram), and confirmed LAEs (red line-filled histogram).}

\caption{Histogram of the broadband (top: EIS-R band and bottom: MUSYC-R band) to narrowband flux density ratio for our z $\sim$ 4.5 LAE sample. The rest-frame \lya\line EW (IGM-corrected\lya\line, marked on the top of the plot) is a monotonic decreasing function of the flux density ratio. The black, blue and green histograms plot distributions for the photometric, targeted, and spectroscopically confirmed samples, respectively. The red line presents the best fit artificial sample assuming the line EW distribution follows an exponential law$d$N$/d$EW = N $exp^{-EW/W_0}$. The best-fit EW exponential scale is $W_0$ = 167$^{+44}_{-19}$ \AA\from MUSYC-R data alone. The horizontal lines are corresponding to the one sigma error on R band flux divided by the minimum narrowband flux.}

\caption{ The EW distribution scale {\color{black}(Bottom: exponential distribution scale $W_0$; Top: Gaussian distribution scale $\sigma_g$) obtained at different redshifts. The diamonds and squares present direct fitting results by assuming exponential and gaussian distributions, respectively, and circles present our simulation approach with two kind of distributions. We also mark results with/without IGM absorption-correction to \lya\line flux in blue/red colors, and we use IGM absorption-correction by Madau (1995) assuming intrinsic symmetric\lya\emission line with zero velocity offset.} Data references: Cowie et al. (2011) at z $\sim$ 0.3, Ciardullo et al. (2012) at z = 2.1 and z = 3.1; Nilsson et al. (2009) at z = 2.25 ;Ouchi et al. (2008) at z = 3.1, and z = 3.7, this work at z = 4.5, Kashikawa et al. (2011) at z=5.7 and 6.5, and Hu et al. (2010) at z = 5.7 and 6.5. {\color{black}The results at z $\sim$ 0.3 are from EGS field sample after AGN have been excluded. } %The green line is the scaled global $f^{Ly\alpha}_{esc}$ from Hayes et al. (2011). The purple line is the scaled dust extinction 10$^{-0.4E_{B-V} k_{Ly\alpha}}$ over the redshift range of 0.3--6.5. }

\caption{Two dimensional distribution of \lya\line flux and broadband-over-narrowband ratio for our observational LAEs and simulated sources. Green plus and cyan squares are the candidates and targets, and red stars are the confirmed LAEs (Bottom: EIS-R\& NB data; Top: MUSYC-R\& NB data, one confirmed LAEs located outside the simulation region is due to the bright neighbor object in the R-band image). The grey points are the simulations with one parameter of exponential scale W$_0$ = {\color{black}167} \AA\on both images, black contours show 50\%, 90\%, and 99\% included regions, and the distribution of the peak value and FWHM of band-ratio distributions as a function of \lya\flux in simulation are plotted as the dashed blue line and the dark yellow region. On the right panel the distributions of\lya\flux for observation (green dashed, all candidates) and simulation (black solid) are shown for EIS-R\& NB data (bottom) and MUSYC-R\& NB data (top), respectively.}

\caption{The rest \lya\EW vs. UV magnitude (the "Ando" effect) for our sample compared with our simulation for EIS-R\& NB (left) and MUSYC-R\& NB (right). Green plus and cyan squares are the candidates and targets, and red stars are the confirmed LAEs. Note that in this plot, we ignore the LAEs with continuum non-detections.}

\caption{Cumulated number density as a function of R band and narrowband limits (5 $\sigma$ limit). The narrowband depth are marked in 4 steps from m$_{lim}$(NB) = 24.5 to 26 presented by different colors. Simulation with $W_0$ = 30, 50, 150, and 300 \AA\are presented as dash-dot, dash, long-dash and solid lines, respectively. The purple filled circles show the narrowband and broad band depth for our narrowband images with EIS-R or MUSYC-R band data, and the brown filled circle shows the depth of LALA Cetus field (Wang et al. 2009).}

136 — 1310.4923

\caption{\label{fig_eta} (Color online) Left panel: in the inset, $\eta$ parameter as a function of $N/\xi$ for various gate voltages ($V_g=1.9\,t$~($\circ$), $2.35\,t$~({\tiny{\color{blue}$\square$}}) and $2.5\,t$~({\color{violet}$\diamond$})), at $E_F=0$ and $W/t=1$. The horizontal lines show the convergence of $\eta$ at large $N$. The asymptotic values are plotted in the main panel as a function of $V_g$. Right panel: $\eta$ parameter in the limit of large $N$ as a function of $V_g$ and $W$, at $E_F=0$. Upon shifting the spectrum of the nanowire with $V_g$, the thermopower distribution moves from a Lorentzian distribution for $V_g\lesssim V_g^c$ ($\eta\approx 1$, blue) to a Gaussian distribution for $V_g\gtrsim V_g^c$ ($\eta\approx 0$, red), where $V_g^c=1.92t+0.34W$ (dashed line).}

137 — 1310.4986

\caption{Enumerating the \dungpreferred{} extensions of an \AFname}

138 — 1310.6160

\caption{Hydrogen concentration profile from Bliersbach measurements as listed in table~\ref{tab:Hydrogen-concentration-profile}, compared with the two other approaches proposed for assessing the diffusion problem (Fick's equation, and computer simulation using OpenFOAM\textregistered{} open-source code).\label{fig:Hydrogen-concentration-profile} }

\caption{The basic OpenFOAM\textregistered{} case structure.\label{tab:Case-struct}}

139 — 1310.7008

\caption{\label{GDmeasurement} (Color online) (a) Raw experimental GD measurements for 1.3~$\mu$m~(\textcolor{r}{$\blacktriangle$}) and 2~$\mu$m~(\textcolor{b}{$\blacktriangledown$}). Estimated GD contributions (offset for clarity) from the Ne atomic phase delay~(\textcolor{c}{\protect\dotdashrule}), dispersion from 0.2~$\mu$m thick Al filter~(\textcolor{g}{\protect\dashedrule}), and neutral Ar dispersion for a gas density-length product of $2.5 \times 10^{17}\, \text{cm}^{-2}$~(\textcolor{m}{\protect\dottedrule}). (b) TDSE-MWE calculations of the total GD~(\textcolor{b}{\protect\mybullet}) for 1.3~$\mu$m, along with the individual $s$-~(\textcolor{g}{\protect\mylozenge}) and $d$-contributions~(\textcolor{r}{\protect\mysquare}) to the GD. In both plots, the attochirp GD~(\protect\solidrule) is shown for reference. Note that the measured macroscopic GD at 1.3~$\mu$m and 2~$\mu$m differ since they have different attochirp GD and were obtained at different medium lengths (2.5~mm and 0.25~mm respectively), thus the Ar dispersion is more apparent in the longer cell 1.3~$\mu$m results. The numbered regions are discussed in the text. }

\caption{\label{dipoleGD} (Color online) Argon measurement~(\textcolor{b}{\protect\mybullet}) and TDSE-MWE calculation~(\textcolor{r}{\protect\solidrule}) of the RDM (a) group delay and (b) phase. (b) also plots the scattering-phase difference $\xi$~(\textcolor{g}{\protect\dashedrule}) used in the simple model, Eq.~(\protect\ref{model_eq}), where the instantaneous $\pi$ jump has been added in by hand. }

\caption{\label{SimpleModel} (Color online) Dipole phase (a) and GD (b) calculated from the simple model in Eq.~(\protect\ref{model_eq}) when using $\Delta\omega=2.2$~eV~(\textcolor{r}{\protect\dashedrule}), 6.6~eV~(\textcolor{b}{\protect\solidrule}), and 13.2~eV~(\textcolor{g}{\protect\dottedrule}). }

140 — 1310.7805

\caption{Hadronic modeling of the \gray flaring spectrum detected by AGILE assuming different indices of the initial proton energy distribution. The blue, red and black solid lines correspond to power-law indices $\alpha=2.4$, $2.5$ and $2.7$, respectively; the dashed line corresponds to $\alpha=2.3$.}

\caption{Panel (a): Calculated \gray unabsorbed (dashed lines) and absorbed (solid lines) emission shown with the \gray flaring flux ($\sim$8 days) detected by AGILE \citep{piano_12}, and the MAGIC upper limits obtained when \cyg was in the soft state (four observation cycles, corresponding to a total of 30.8 hours, \citealp{aleksic_10}). The \gray absorption by UV stellar photons is calculated for different distances $r$ from the companion star. In panel (b) the corresponding calculated spectrum of HE neutrinos is presented, the filled area represents the maximum flux of HE neutrinos. In panel (b) the solid blue line corresponds to a neutrino model with the same parameters of the blue dot-dashed curve ($r=R_\star$, $\alpha=2.4$), but with a maximum energy of protons of $10\:\rm{PeV}$. Red and blue dashed lines in panel (b) corresponds to IceCube sensitivities expected for two-month (61 days) and one-year (365 days) exposure time, respectively.}

141 — 1310.8515

\caption{The light curve of Asteroid (596) Scheila before (black and white, \citet{Warner:vz}) and after the impact (blue symbols, \citet{Ishiguro:2011to}). Different symbols indicate different observing days. Before impact: \textcolor[gray]{0.7}{$\blacklozenge$} - 1/10/2006; $\circ$ - 1/13/2006; \textcolor[gray]{0.7}{$\blacktriangle$} - 1/21/2006; $\vartriangle$ - 1/22/2006; $\bullet$ - 1/23/2006; $\blacksquare$ - 1/24/2006. Post impact: \textcolor{blue}{$\blacklozenge$} - 12/17/2010; \textcolor{blue}{$\circ$} - 12/20/2010; \textcolor{blue}{$\vartriangle$} - 12/23/2010; \textcolor{blue}{$\blacktriangle$} - 12/26/2010; \textcolor{blue}{$\ast$} - 1/2/2011; \textcolor{blue}{$\square$} - 1/3/2011; \textcolor{blue}{$\blacksquare$} - 1/7/2011; \textcolor{blue}{$\bullet$} - 1/8/2011; \textcolor{blue}{$\lozenge$} - 1/28/2011. The dashed vertical lines indicate the part of the lightcurve that was affected most by the impact. }

\caption{Subset of the light curve measurements of Asteroid (596) Scheila measured at phase angles of approximately 11 degrees. Symbols are equal to those used in Fig. 1; Different symbols indicate different observing days. Before impact: \textcolor[gray]{0.7}{$\blacklozenge$} - $\phi$ = 10.7 deg; $\circ$ - $\phi$ = 9.97 deg; Post impact: \textcolor{blue}{$\blacksquare$} - $\phi$ = 11.26 deg; \textcolor{blue}{$\bullet$} - $\phi$ = 11.02 deg. The dashed vertical lines indicate the part of the lightcurve that was affected most by the impact. }

142 — 1311.0456

\caption{\Shared\full-duplex\label{fig:shared-antenna}}

143 — 1311.0910

\caption{ (color online). Nine particle tracks which highlight the characteristic features of each injection mechanism. Three LIDA tracks are plotted in red, blue, and green in \textcolor{blue}{a} and \textcolor{blue}{d}; three laser ionization-injection tracks are plotted in black, pink, and cyan in \textcolor{blue}{b} and \textcolor{blue}{e}; three other tracks are plotted in dark green, orange, and gray in \textcolor{blue}{c} and \textcolor{blue}{f}. For each case, the top row plots the kinetic energy (signed by $p_{z}$) vs. longitudinal position (\textcolor{blue}{a}, \textcolor{blue}{b}, \textcolor{blue}{c}) and the bottom row plots the trajectory (\textcolor{blue}{d}, \textcolor{blue}{e}, \textcolor{blue}{f}). In \textcolor{blue}{d}, \textcolor{blue}{e}, and \textcolor{blue}{f} the location of the critical density is shown by the dashed line. The $x$-dependent relativistic critical density is calculated using the peak field at each point in $x$ for the initial plasma conditions.}

144 — 1311.1361

\caption{(Color online) Electronic band structures of (a): LaSn$_3$ and (b): YSn$_3$. The solid lines (red colour) show the electronic levels calculated with spin-orbit coupling included, while the dotted lines (blue colour) show the electronic levels as calculated without spin-orbit coupling. The energies are given in eV relative to the Fermi level, $E_F$, which is marked with the horizontal dashed line. The major difference between the two compounds around the Fermi level occurs in the vicinity of the X-point (for SOC included). \textcolor{blue}{ \bf A second, less significant feature, is a very dispersive band along $\Gamma-R$, which stays above $E_F$ for LaSn$_3$, but dips below $E_F$ for YSn$_3$. The inset illustrates this.}}

145 — 1311.1714

\caption{An example graph and its representation in the graph format. The IDs of the vertices are drawn within the cycle, the vertex weight is shown next to the circle ({\color{red}red}) and the edge weight is plotted next to the edge ({\color{blue}blue}).}

146 — 1311.1776

\caption{(a) Photographs of the target object, a Lego NINJAGO Samukai\textregistered~figure. (b) Typical frame of data containing 96 photons in a 396x266 pixel detector. This translates to \mbox{$9.1\times10^{-4}$} photons/pixel. (c) Data frame with 1025 photons obtained by combining 10 consecutive frames from the previous data set. The sizes of the pixels recording photons have been enlarged to improve visibility. Pixels with two photons are shown in red.}

147 — 1311.1939

\caption{Success rate (SR)(\%). \textcolor{red}{\textbf{Red}} fonts indicate the best performance while the \textcolor{blue}{\textbf{blue}} fonts indicate the second best ones. The total number of evaluated frames is $7,591$. % (Sort in alphabetical order). }

\caption{Center location error (CLE)(in pixels) and average frame per second (FPS). \textcolor{red}{\textbf{Red}} fonts indicate the best performance while the \textcolor{blue}{\textbf{blue}} fonts indicate the second best ones. The total number of evaluated frames is $7,591$. % (Sort in alphabetical order). }

148 — 1311.4144

\caption{Lepton - lepton separations for the same sign lepton pairs (\textcolor{blue}{$\Delta R_{\ell^{\pm} \ell^{\pm}}$}) and opposite sign lepton pairs (\textcolor{red}{$\Delta R_{\ell^{\pm} \ell^{\mp}}$}) for $(pp\rightarrow H_{1,2}^{++} H_{1,2}^{--} \to 4 l)$ within the degenerate scenario with $M_{H_1^{++}} = M_{H_2^{++}} = 400$ GeV for $\sqrt{s}=14$ TeV and $L=300 ~fb^{-1}$. }

\caption{Lepton - lepton separations for same sign lepton pairs (\textcolor{blue}{$\Delta R_{\ell^{\pm} \ell^{\pm}}$}) and opposite sign lepton pairs (\textcolor{red}{$\Delta R_{\ell^{\pm} \ell^{\mp}}$}) in the $(pp\rightarrow H_{1,2}^{++} H_{1,2}^{--} \to 4 l)$ process for non-degenerate mass scenario having $M_{H_1^{\pm\pm}} = 400$ GeV and $M_{H_2^{\pm\pm}} = 500$ GeV with $\sqrt{s}=14$ TeV and $L=300 ~fb^{-1}$. }

\caption{Lepton-lepton separation plot for same sign leptons (\textcolor{blue}{$\Delta R_{\ell^{\pm} \ell^{\pm}}$}) and opposite sign leptons (\textcolor{red}{$\Delta R_{\ell^{\pm} \ell^{\mp}}$}) in the process $(pp\rightarrow (H^{\pm \pm}_{1} H^{\mp}_{1} + H^{\pm \pm}_{2} H^{\mp}_{2}) \rightarrow 3 \ell )$. Here $\sqrt{s} = 14$ TeV and integrated luminosity 300 $fb^{-1}$ at the LHC. }

149 — 1311.4346

\caption{Phase diagram of the three-spin interacting transverse Ising model along with the various paths studied for approaching the MCP. \textcolor{blue}{The point A corresponds to one of the multicritical points. The phase boundaries are marked by the three different lines as shown in the label whereas the paths studied in this paper are I, II, III and IV, as shown by the lines with arrows. Path IV is also the gapless line separating various phases. The shaded region corresponds to the region where quasicritical points exists. }}

150 — 1311.4699

\caption{ \label{figs:ratios} Left: plot of the ratio $R$ [see Eq.\(\protect\ref{eq:ratio})], as a function of the iteration step $i$, for a typical configuration and lattice volume $V=16^4$. Right: plot of $\lambda_2$ (\textcolor{red}{\bf full circes}), $| {\cal E}''' \, |$ (\textcolor{cyan}{\bf full squares}) and $ {\cal E}'''' \, $ (\textcolor{blue}{\bf full triangles}) as a function of the iteration step $i$, for the same configuration considered in the left plot. }

\caption{\label{fig:ineqnew} Left: plot of the (normalized) horizon function $h$ (\textcolor{blue}{\bf empty circles}), of the Gribov ghost form-factor $\sigma(p_{min})$ ({\bf full triangles}), of the quantity $1 - \lambda_1/p_{min}^2$ (\textcolor{cyan}{\bf full squares}) and of their upper bound $\rho$ (\textcolor{red}{\bf full circles}) as a function of the inverse lattice size $1/N$. Let us note that, in Ref.\\cite{Capri:2012wx}, it was proven that $\sigma(0) = h$ to all orders in the gauge coupling. Right: plot of the inverse of the lower bound in Eq.\(\protect\ref{eq:lowerb}) (\textcolor{blue}{\bf empty circles}), of $1/G(p_{min})$ ({\bf full triangles}), of $\lambda_1$ (\textcolor{cyan}{\bf full squares}) and of the quantity $(1 - \rho) \, p^2_{min}$ (\textcolor{red}{\bf full circles}) as a function of the inverse lattice size $1/N$. }

151 — 1311.5201

\caption{(Color online) Orientation dependence of the absorbed energy $E_\text{abs}$ normalized to its maximum value $E_\text{abs}^\text{max}$ (given in each figure) in a linearly polarized laser field {$\marker{\mathbf{E}(t;\theta,\phi)}$} with different laser parameters ($T = 10 \text{\ fs}$ in all cases): \\ - left column (a),(c): $\lambda = 800\ \text{nm}$ (non-resonant) \\ - right column (b),(d): $\lambda = 400\ \text{nm}$ (resonant) \\ - first row (a),(b): $I = 1.0 \cdot 10^{13} \frac{\text{W}}{\text{cm}^2}$ \\ - second row (c),(d): $I = 1.2 \cdot 10^{14} \frac{\text{W}}{\text{cm}^2}$ \\ The black dots indicate the positions of the carbon atoms {\marker{(view on yz-plane, cf. sketch on the top of the Figure, where we have slightly rotated the molecule for a complete definition of the polarization respectively orientation angles)}}. The results have been calculated using the fixed-nuclei approximation.}

\caption{(Color online) $\text{C}_{60}$ radius $R$ as a function of time $t$ for different molecular orientations (dotted lines) in the same laser field ($T = 54 \text{\ fs}$ , $I = 6.2 \cdot 10^{14} \frac{\text{W}}{\text{cm}^2}$, $\lambda = 800 \ \text{nm}$) as used before (cf. Fig. \ref{fig:long}). The colors (blue, green, red) correspond to typical orientations within the same colored regions of Fig. \ref{fig:long} (b) and, thus, belong to different energy absorption intervals given {\marker{in the Figure}}. The full lines are the corresponding mean values of $R(t)$ and summarize the three different relaxation channels: (1) excitation of breathing modes (blue), (2) formation of non-breathing isomeric states (green), and (3) ultrafast almost symmetric fission (red). {\marker{Typical snapshots of the dynamics at $t \approx 240 \text{\ fs}$ corresponding to the three relaxation channels are shown on the top of the Figure using the same colors. The black arrows indicate the laser polarization axis [cf. Fig. \ref{fig:long} (b)]. Note, the fission process proceeds along the laser polarization direction.}}}

152 — 1311.5205

\caption{\label{fig:onset-PG}The values of $T_{c}^{\prime}$ (\fullsquare) and $T^{\star}$ (\fullcircle) are plotted versus the hole doping for the four samples studied (from ref.\cite{FRA-PRB2011}). The solid line indicates the superconducting dome. Contrary to $T_{c}^{\prime}$ that is rather insensitive to hole doping, $T^{\star}$ is found to decrease with increasing doping and intersects the $T_{c}^{\prime}$ line near optimal doping \cite{Alloul-EPL2010}.}

\caption{\label{fig:RMN-SCF}The fluctuation conductivity (\fullcircle, left scale) is compared to the Y NMR Knight shift \cite{Alloul-1989}, \cite{Alloul-1993} (\fullsquare, right scale) for the UD57 sample. It is remarkable to see that nearly half of the susceptibility has already been lost at the onset of SCFs.}

153 — 1311.5223

\caption{Percentage differences between WDM and CDM non-linear matter power spectra from hydro-dynamical simulations at high-resolution. The solid lines show the large scale power, while the dashed lines describe the small scale power obtained with the folding method in order to reach smaller scales \citep[see][for details]{Jenkins:1998,Colombi:2009}. The dotted line is the suppression to the linear matter power spectrum and is the same both in the $z=3$ and $z=0$ panels. The different panels show $z=0,0.5,1.2,3$. \redtxt{Note that the steep rise on scales, $k>50 \, h/Mpc$ is affected by the poor resolution of the WDM simulations and it is not fully physical (although an increase of power could be expected and it might be due to the different dark matter density profile at small scales).} }

\caption{ \redtxt{On the left we plot the mass functions from \cite{Sheth:1999} (dotted) vs. \citet{Schneider:2012} fit to simulations (solid). On the right we show the re-scaled ratios of these mass functions by the half-mode halo mass, $M_{\rm hm}$.} The half-mode and the free-streaming halo masses are plotted with asterisks and crosses, respectively. }

\caption{NFW \citep{Navarro:1997} halo density profiles for two different halo masses ($10^{10}$ and $10^{12}$ $M_\odot$) calculated with the re-scaled concentration parameter from \citet{Schneider:2012} for WDM models with $m_\WDM~\in~\{0.25,0.5,1.0,2.0,10.0\}\keV$. \redtxt{The dashed vertical lines correspond to the free-streaming lengths in the different WDM models, rescaled by the virial radius of the halo.} }

154 — 1311.6184

\caption{The estimates of the log-probabilities of the test samples for 12 RBMs with varying numbers of latent variables. The curves represent the log-probabilities estimated using AIS (\textcolor{blue}{blue}), the biased CSL with a single step of 10 parallel Markov chains (\textcolor{red}{red}), the biased CSL with 300 steps of 10 parallel Markov chains (\textcolor{cyan}{cyan}) and the true log-probabilities (only for the small models, \textcolor{green}{green}). }

155 — 1311.6567

\caption{\label{fig-1} \tred{Normalized mean square error} of $\hat{\Gr M}$ estimated by the FPE and the shrinkage FPE versus the parameter $\beta$ of the shrinkage FPE and for different covariance matrices, i.e. for different values of the correlation coefficient $\rho$, where $N=24$ samples and the dimension of the data is $m=12$ for Gaussian noise.}

156 — 1311.7632

\caption{\label{fig:psslc-central-tori}% Central lines \Ca, \Cb, \Cpo, \Cres{} (\colorcentera, \colorcenterb, \colorperiodseven, \colorthreetower) in the \psslc{} representing families of $1$-tori of type \alp{} \Ma, \Mb, \Mpo, and type \bet{} \Mres{} for two coupled standard maps, Eq.~(\ref{eq:CoupledStdMaps}). Around the central lines the $2$-tori (\colortwotorus) are organized. From the elliptic-elliptic fixed point \fixedpoint{} the two central lines \Ca, \Cb{} (\colorcentera, \colorcenterb) emanate, and continue beyond the large gaps. The central lines \Cres{} (\colorthreetower{}) arise from a $-1:3:0$ resonance. \MOVIEREF}

\caption{\label{fig:fa-central-tori}% Frequencies $(\nu_1, \nu_2)$ of all $2$-tori (grey). The largest gaps stem from the $3:1:1$ and $-1:2:0$ resonances. The frequencies of the fixed point \fixedpoint{} are $(\valfixpt{1}, \valfixpt{2})=(0.30632, 0.12173)$. From there the \alp{} families of $1$-tori \Ma, \Mb{} (\colorcentera{}, \colorcenterb{}) emanate, forming the skeleton of the regular region. The frequencies for $1$-tori of the \bet{} family \Mres{} (\colorresonancethreetower{}) are all on the $-1:3:0$ resonance line and for the \alp{} families \Mpo{} (\colorperiodseven{}) on the point $(\nicefrac{2}{7}, \nicefrac{1}{7})$. The families \Mh{\alpm{2}}{} (\coloralphab) are discussed in Sec.~\ref{sec:hierarchy}.}

\caption{\label{fig:hierarchy}% First level of the hierarchy for \subfiga{} type \alp{1}, \subfigb{} type \bet{}, and \subfigc{} type \alp{2}: Magnified views of Fig.~\ref{fig:psslc-central-tori} for \subfiga{} \Cpo{} around one point $\periodicpoint=(0.0, 0.0, 0.083438087, 0.118666288)$ of the elliptic-elliptic period-$7$ orbit, \subfigb{} a branch of \Cres{} of the $-1:3:0$ resonance, and \subfigc{} central lines \Caa{1} (\coloralphaa) and \Caa{2} (\coloralphab) representing families of $1$-tori \Maa{1} and \Maa{2} around one point $\alphapoint=(0.115287658, -0.141621338, 0.0, 0.0)$ of the elliptic-elliptic period-$7$ chain. The slice condition is for \subfiga{}, \subfigb{} $p_2^*=0$ and $\sectioneps=10^{-5}$ and for \subfigc{} $q_2^*=0$ and $\sectioneps=10^{-6}$. The coordinate system in \subfiga{} and \subfigc{} is moved to \periodicpoint{} and \alphapoint{} respectively, and in \subfigb{} to an arbitrary point along \Cres. In \subfigd{}, \subfige{}, and \subfigf{} the adapted frequencies are shown. In \subfiga{} and \subfigd{} the periodic point \periodicpoint{} with frequencies $(\nu_1', \nu_2')=(0.1515, 0.0838)$ is marked by a magenta point. The two emanating families of $1$-tori are colored the same for simplicity and have edges intersecting in the frequency plane with an angle close to $\pi$. In \subfigc{} and \subfigf{} the periodic point \alphapoint{} with frequencies $(\nu_1''', \nu_2''')=(0.4475, 0.0081)$ is marked by a red point. The inset in \subfigf{} shows a magnification around \alphapoint. \MOVIEREF}

157 — 1312.0159

\caption{(Color online) Switching field $B_{c}$ extracted from anisotropic magnetoresistance (AMR) measurements on individual Py strips as a function of the strip width $w$. All strips are $25\,$nm thick and $10\,\mu$m long. The symbols represent values for strips obtained by \textcolor{red}{$\bullet$}~thermal evaporation and ZEP recipe, \textcolor{green}{$\blacktriangle$}~sputter deposition and ZEP masks, and \textcolor{black}{$\blacksquare$}~thermal evaporation and an optimised PMMA/MA recipe. The dashed line is a guide to the eye. The inset shows AMR curves of a $w=180\,$nm sputtered Py strip.}

158 — 1312.0579

\caption{\StructuredSpeedBoost}

159 — 1312.0957

\caption[m]{Polarizability versus length to width ratio (\begin{tikzpicture}\draw[red] (0,0)--(0.6,0); \draw[red] (0.2,-0.1) rectangle (0.4,0.1); \end{tikzpicture} box containing antenna, \tikz \draw[blue,thick] (0,0) circle [radius=0.075]; meander line antenna, \tikz \draw[thick] (0,0.1)--(0.5,0.1); rectangle)}

160 — 1312.2026

\caption[Unsuccessful Stokes-I fits]{Both of these Stokes-I fits belong to pixels whose reported B-value is a clear outlier (too high) from where it should be (the one in the second panel being a less-extreme outlier than the one in the first) Note, however, \textcolor{red}{that for PROVER2B=1.2 and later} that {\em all} lines fits are now {\em forbidden} to be in emission.}

161 — 1312.2399

\caption{Integrated (35-arcsec-thick) major-axis position-velocity slices extracted from the NGC~3521 \hi\data cubes. Panel~\textit{a}: the full THINGS cube containing all of the \hi\emission. Panel~\textit{b}: the regular \hi\data cube containing emission from the thin\hi\disc. Panel~\textit{c}: the anomalous \hi\data cube containing emission from the slow-rotating\hi\mass component. Identical colour scales are used for panels$a$ and $b$ while a separate scale is used for panel~$c$. \hi\column densities to which the colours in each panel correspond are given by the colour bars in units of$10^{20}$~cm$^{-2}$. Note the much lower column densities of the anomalous \hi\emission in panel$c$.}

\caption{{\color{black}Data representing a typical \hi\line profile (black), fitted by a double Gaussian (grey). The dashed horizontal line marks the$2\sigma$ noise level of the data. The hatched region of each Gaussian component represents 20~per~cent of its total area, measured vertically from the peak of the profile towards zero. The fitted component on the right is accepted, while the component on the left is rejected.}}

\caption{Line profiles extracted from the THINGS \hi\data cube. In each panel the solid black curve represents the line profile. The thick grey curve represents the double-component Gaussian fit. The blue-dashed and red-dotted curves represent the Gaussian components corresponding to the regular and anomalous\hi, respectively. The vertical and horizontal dashed lines mark the systemic velocity of the galaxy and the $2\sigma$ noise level (7.8~mJy~beam$^{-1}$) of the noise in a line-free channel of the cube, respectively.}

\caption{Global profiles extracted from the decomposed \hi\data cubes of NGC~3521. The blue long-dashed and red short-dashed curves represent the total flux within each channel of the regular and anomalous\hi\cubes, respectively. The solid black curve represents the global profile of the full THINGS data cube (from which the regular and anomalous\hi\data cubes were extracted).}

\caption{Azimuthaly-averaged \hi\column densities for the regular (blue long-dashed) and anomalous (red short-dashed)\hi\components in NGC~3521.}

\caption{NGC~3521 \hi\velocity fields. Panel~\textit{a/b}: Velocity field of the regular/anomalous \hi\component of the galaxy. In each panel the iso-velocity contours are separated by 25~\kms. The thick iso-velocity contour at 798.2~\kms\marks the systemic velocity of the galaxy. Both panels use the same colour scale, given in units of\kms\by the colour bar. The hatched circle in the lower right corner of panel\textit{a} represents the half power beam width of the synthesised beam. The scale bar corresponds to angular and physical lengths of 150~arcsec and 7.7~kpc, respectively. Panel~\textit{c}: residual velocities obtained by subtracting the anomalous \hi\velocity field from the regular\hi\velocity field. Units are given in\kms\by the colour bar.{\color{black}The red-dotted contours overlaid on the southern half of the galaxy are the same as those shown in the channel map at $V_{hel}=716.9$~\kms.}}

\caption{Panels $a,c:$ Residual velocity fields, generated by subtracting the tilted ring models from the data. Units are given in \kms\by the colour bars. Panels$b,d:$ Ratio of the absolute residual velocities to the corresponding rotation velocities.}

\caption{Position-velocity slices extracted from the \hi\data cube and the model cube. Each of rows 1~-~3 show three different slices extracted parallel, perpendicular and at 45~degrees to the major axis of the\hi\disc, respectively. The slices are shown as red lines in the\hi\total intensity maps. Each slice is presented as a data-model pair. The\hi\total intensity map is taken from the THINGS pubic data repository (http://www.mpia-hd.mpg.de/THINGS/Data.html). The white ellipse shown in panel$1d$ roughly delimits the velocity gap between the regular and anomalous \hi\components of the model cube. For the model cube presented here, the anomalous\hi\is treated as being distributed in a thick ($\sim 3.5$~kpc), slow-rotating disc with no vertical gradient in the direction perpendicular to the \hi\disc. No thin-disc models for the anomalous\hi\can match the observations.}

\caption{{\color{black}Position-velocity slice extracted parallel to the major axis of the galaxy in the de-rotated \hi\cube. More precisely, major axis position-velocity slices were extracted at each point on the minor axis and summed together to produce the image shown here. The broad band of de-rotated\hi\emission at$V_{dev}=0$~\kms\corresponds to the thin\hi\disc in NGC~3521. The remaining emission in the range -100~\kms~$\lesssim~V_{dev}~\lesssim$~100~\kms\is the anomalous\hi. The white curve represents the strip-integrated 12~\micron\emission (in arbitrary units). The vertical dashed lines mark the radius containing 90~per~cent of the 12~\micron\emission;$R_{90}=189$~arcsec. Clearly evident is the fact that the vast majority of anomalous \hi\in NGC~3521 is spatially coincident with the inner regions of the disc where the star formation is highest.}}

162 — 1312.2845

\caption{\rodcolor The measurement of field-effect dependence of thermoelectric effects in single nanostructures (a nanowire device is visible in panel (a)) requires the application of a strong thermal gradient (panel (b)). A standard approach consists in the fabrication of a top heating element (panel (c)). An alternative ``buried'' architecture exploiting current flows into the bulk is visible in panel (d).}

\caption{\rodcolor (a) Scanning electron micrograph of one of the fabricated buried heaters. (b) Sketch of the contact geometry: the heater is designed to achieve the best differential heating conditions in the red circle are, which is the one where the nanostructure is supposed to reside. (c) Blow-up image of one of the current injection contacts used to feed current to the substrate through a hole in the $280\,{\rm nm}$-oxide covering the Si substrate (sample was tilted by $60$ degrees) . (d) Cross-section sketch of the contact region.}

\caption{\rodcolor Finite element estimate of the temperature profiles for different values of the $L/W$ ratio in the studied geometry and assuming an ideal thermal sinking at the heater contacts for sake of simplicity. The colorplot reports the temperature increase $\Delta T$ as a per cent of the maximum value in the heated regions. For a fixed $L=10\,{\rm \mu m}$ the (a), (b) and (c) panels report the case $W=4$, $9$ and $14\,{\rm \mu m}$, respectively. The red contour lines corresponds to the highest gradient along the symmetry line of the heater while the dashed lines are spaced by $5\%$ of the top $\Delta T$. An optimal uniformity at the top gradient position is achieved for $H\approx L$.}

\caption{(a) IV characteristics of the heater before (blue), during (magenta) and after (red) the self-annealing procedure. Top inset: a high driving current can lead to a time-dependent evolution of the IV. Bottom insets: the drift of the heater voltage $V_H$ at fixed $I_H$ and as a function of time. (b,c) Voltage difference between the two substrate probes $V_{P1}$-$V_{P2}$ and individual value of $V_{P2}$. {\rodcolor Both panels report the characteristics ad various stages of the self-annealing procedure, according the the plot color.}}

\caption{\rodcolor (a) Raman map of the local temperature increase of the SiO$_2$/Si substrate in the vicinity of the heater and of the device measurement site. (b) Numerical simulation for a injected power of about $300\,{\rm mW}$ including the lead finite heat conductivity and contact dissipation. (c) The measured temperature profile along the red dashed line on the Raman map. (d) Temperature difference between two resistive thermometers spaced $1\,{\rm \mu m}$ (see Fig.~\ref{fig:SemDevice}a) as a function of the electrical power fed heater. The green arrows highlight the power used for the Raman map.}

\caption{\rodcolor (a) Scanning electron micrograph of the studied nanowire device and (b) blow-up of the isolated nanostructure. (c) Measurement of the field-effect evolution of the Seebeck coefficient plotted parametrically against the nanowire resistance. The application of a strong gradient of $\approx 5\,{\rm K/\mu m}$ allows to obtain a much clearer evidence of the effect.}

163 — 1312.3335

\caption{Measured He\,\textsc{i} radial velocities for SDSS\,J1137 folded on a period of 59.6\,minutes. The dashed and dotted lines are the best fitting radial velocity curve and systemic velocity.\label{f:RV1137}}

164 — 1312.4130

\caption{(a) Speed $v$ as a function of concentration of CTAB ($\blacktriangle$), KBr ( \tikzcircle{2pt}) , NaCl($\circ$), and NaOH (\textcolor{blue}{$\blacksquare$}). Positive $v$ indicates propulsion towards PS. Leftmost points correspond to zero concentration. (b) Translational diffusion $D$ for the same data. The dotted horizontal line corresponds to the predicted bulk diffusion constant $D=k_BT/(6\pi\eta R)=0.23\mathrm{~\mu m^2s^{-1}}$.}

\caption{Fractional mass loss of \ce{H2O2} over time from $10^{-5}$ v/v Janus particles in $10\%$ \ce{H2O2}, with ($\circ$) or without (\textcolor{red}{$\triangle$}) 1~mM~NaCl. The control (\textcolor{blue}{$\square$}) is $10\%$ \ce{H2O2} without Janus particles. Solid lines are exponential fits to the data.}

\caption{Speed against inverse solution conductivity for KBr ( \tikzcircle{2pt}) and NaCl($\circ$), from Fig.~\ref{All figures plot}. The solid line is a linear fit to the data, described in the text. }

165 — 1312.4360

\caption{The ratio between the NNN quantum tunneling and the NN quantum tunneling ($\delta E/\delta E^{\prime}$) via flux-superlattice constant $L$, where $\delta E$ versus $L$ is shown in Fig.\textcolor[rgb]{1.00,0.00,0.50}{4}. The inset shows triangular flux-superlattice. The tunnelings $\delta E$, $\delta E^{\prime}$ and the flux-superlattice constant $L$ are all indicated in the inset. }

166 — 1312.4512

\caption{Dimensionless elastic modulus $\hat{G}$ as a function of the gas volume fraction $\phi$ for dispersions with three different bubble radii $R_b$ in emulsion (3) [see legend]. The full lines are the computed $\hat{G}_{homog}(\phi)$ for $Ca=0.23$ (dark blue), $Ca=0.57$ (green) and $Ca=3.2$ (pink); experimentally measured $Ca$: $0.23\pm0.05$, $0.57\pm0.08$, $3.2\pm0.4$. Inset: $\hat{G}$ as a function of $\phi$ for dispersions of $R_b\approx 150\mathrm{\mu m}$ in \textcolor{spinach}{$\bullet$} emulsion (3) and \textcolor{orange}{$\blacklozenge$} emulsion (4). The full lines are the computed $\hat{G}_{homog}(\phi)$ for $Ca=0.57$ (green) and $Ca=1.65$ (orange); experimentally measured $Ca$: $0.57\pm0.08$, $1.65\pm0.15$.\label{fig:1} }

\caption{Effect of a change in the surface tension: dimensionless elastic modulus $\hat{G}$ as a function of $\phi$ for dispersions of \textcolor{spinach}{$\bullet$} $R_b=143\mathrm{\mu m}$ bubbles in emulsion (3) and \textcolor{red}{$\bigstar$} $R_b=129\mathrm{\mu m}$ bubbles in emulsion (1). The surface tension is much lower in emulsion (1b), but $G'(0)$ has been chosen to get close values for $Ca$ in both systems. This experimentally measured $Ca$ are $0.57\pm0.08$ and $0.70\pm0.08$. The full line is the computed $\hat{G}_{homog}(\phi, Ca)$ at $Ca=0.63$, which is compatible with both systems, given the uncertainty on the value of $Ca$.\label{fig:2}}

\caption{Remarkable values of $Ca$: $\mathbf{Ca\to \infty}$: $\hat{G}(\phi)$ for dispersions of \textcolor{lgtblue}{$\blacktriangledown$} $R_b$=1mm bubbles in emulsion (2). The full line is the computed $\hat{G}_{homog}(\phi, Ca)$ at $Ca=9.0$ (light blue); experimentally measured $Ca$: $9.0\pm1.2$. $\mathbf{Ca\approx 0.25}$: dimensionless elastic modulus $\hat{G}$ for dispersions of \textcolor{darkblue}{$\bullet$} $R_b$=50$\mu$m bubbles in emulsion (3) and \textcolor{gray}{$\bigstar$} $R_b$=41$\mu$m in emulsion (1a). The full lines are the computed $\hat{G}_{homog}(\phi, Ca)$ at $Ca=0.23$ (dark blue) and $Ca=0.30$ (grey); experimentally measured $Ca$: $0.23\pm0.05$, $0.30\pm0.05$. The dashed lines are the dilute limits for rigid (top) and fully deformable (bottom) spheres, with a full slip boundary condition.\label{fig:3} }

167 — 1312.4729

\caption{\textcolor{red}{(Color online)} Showing the upper tetrahedron and first octahedron of a generation-2 hierarchical structure constructed through a rapid prototyping technique. This structure was created through use of EnvisionTEC Perfactory machine. Inset shows the layering effect of the rapid prototyping procedure. The layer thickness of the structure shown is approximately 25$\mu$m. The material used in the construction of this structure is EnvisionTec R05 \cite{envisionTEC}.\label{structure}}

\caption{\textcolor{red}{(Color online)} The material saving through use of a generation-$G$ hollow tube structure when compared to a solid beam. The plot is valid for a material with Poisson ratio of $\nu \approx 0.3$. The progression of optimality for higher generation designs is clearly shown with the increase in length, $L$, or decrease in force, $F$, for a given Young's Modulus, $Y$. Also depicted are regions showing typical parameters for some compression bearing structures: approximate regions for steel crane booms \cite{liebherr}, iron chair legs, solar sail compression beams \cite{Solar_Sail} (from an arbitrary stiff material, $Y>100$GPa) and mammal femurs withstanding only static loads \cite{Trab_Allo}. Also shown is the positioning of the test problem investigated in the text and in table \ref{default} ($F = 10$kN, $L = 200$m and $Y = 210$GPa). \label{complexity}}

168 — 1312.4766

\caption{Shown are the photo-electron spectra for different peak intensities marked by arrows in the lower panels of Fig. {\color{blue} 2} for the case of $\omega$=40.8 eV (resonance condition for the residual ionic states $|1\textit{s}\rangle$ and $|2\textit{p}\rangle$). The left panels show the main emission line corresponding to the ionic final state $|1\textit{s}\rangle$, the right panels correspond to photoelectron emission with excited ionic final states.}

\caption{ Same as in Fig. {\color{blue} 3}, but for the case of $\omega$=48.356 eV (resonance condition for ionic states $|1\textit{s}\rangle$ and $|3\textit{p}\rangle$).}

169 — 1312.5325

\caption{Plot of the distribution of monojet signal events as a function of invariant mass of the $\bar{\chi}\chi$ and jet $p_T$ in different EFTs. The normalization of the event distribution is arbitrary. \blue{We can see that most events are produced with a large center of mass energy, greater than the current bound on the scale of the effective operators, $M_*$ .}}

170 — 1312.5784

\caption{Crossed-trap lifetimes along the waveguide; {\scriptsize\textcolor{blaa}{$\blacksquare$}}, single channel FSK with $f_X=72.4$~MHz (fixed) optimized for $z=0$; {\large$\bullet$}, dual-channel FSK with $f_X$ varying linearly with $z$ and optimized lifetimes at $z=\pm1.8$~mm. Optimizations were based on the data displayed in the inset: number of atoms left $100$~ms after loading site versus $f_X$ at $z =\{ -1.8, 0, 1.8\}$~mm. Optimized $f_X(z)$ trajectories for single and dual-channel FSK shown as solid blue and dashed black lines, respectively.\label{fig:xtoggle}}

171 — 1312.7377

\caption{The adaptive coupling gains \textcolor{red}{$c_{i}$} in \dref{ssda}. }

172 — 1312.7843

\caption{Degree distribution exponent $\nu$ versus $\lambda$ for enhanced redirection with single attachment (\textcolor{blue}{$\circ$}) and double attachment (\textcolor{red}{$\Delta$}). Each data point is determined from fits of $N_k$ versus $N$, as in Figs.~\ref{NkVsN}(a) and \ref{NkDouble}. }

\caption{Probabilities for a star (\textcolor{blue}{$\circ$}) and a hairball graph (\textcolor{red}{$\Delta$}), $\mathcal{S}$ and $\mathcal{H}$, respectively. For each $\lambda$, the data are based on $10^5$ realizations. The solid curve represents the numerical evaluation of the product in \eqref{S_max}. }