2012

1 — 1201.0165

\caption{High-energy spectrum of thorium wire on HPGe surface. Characteristic steps corresponding to initial kinetic energies of alphas in \iso{232}{Th} decay chains are marked. The lower-energy spectrum is dominated by gammas from the chain, and the lines visible above 3 MeV are double and triple coincidences of the 583 keV, 2615 keV, and 511 keV gammas from \iso{208}{Tl}.{\color{red} yep, i realize it has no label. trying to figure out why the model is'nt replicating the step endpoints well (the model steps are higher in energy and steeper than the data)}.}

2 — 1201.0433

\caption{\label{Fig:EigenvalueSpectrum} (Color online) Eigenvalue spectrum of the correlation matrix of inventory variation of investors trading stock 000001 within 1 day time horizon in 2003. The solid line is the spectral density obtained by shuffling independently the buyers and the sellers in such a way to maintain the same number of purchases and sales for each investor as in the real data. The dashed blue line shows the spectral density predicted by the random matrix theory using Eq.~(\ref{Eq:RMTPDF}) with $Q=237/80=2.96$. The inset shows the largest eigenvalue $\lambda_1$ ($\bigcirc$) and the second largest eigenvalue $\lambda_2$ ({\color{red}{$\square$}}) of all 15 investigated data sets from 15 stocks. The solid line indicates the upper thresholds by shuffling experiments, and the dashed line presents the threshold predicted by the random matrix theory.}

\caption{\label{Fig:Categorization:test} (Color online) Panels (a-c) show the scatter plots of $C_{VR}$ versus a proxy of the size of the investor. For each stock, the proxy is the ratio of the value exchanged by the investor (scaled by a factor $10^4$) to the capitalization of the stock. Each marker refers to an investor trading a specific stock. The three kinds of markers refer to investors whose inventory variations are positively correlated ({\color{red}{$\bigcirc$}}), negatively correlated ({\color{green}{$\square$}}), or uncorrelated ($\vartriangle$) with returns according to the block bootstrap analysis. The two dashed lines indicate the $2\sigma$ threshold calculated using Eq.~(\ref{Eq:threshold}). Panels (d-f) are contour plots of the correlation matrix of daily inventory variation of investors trading the stock 000001. We have sorted the investors into rows and columns according to their cross-correlation coefficients of inventory variation with its price return $C_{VR}$. The evolution of $C_{VR}$ in the same order as in the matrix is shown in the bottom panel, where the dashed lines bound the $\pm 2\sigma$ significance intervals.}

3 — 1201.0663

\caption{\label{tmsgate2} The relativistic motion of the cavity is used to produce a two mode squeezing gate. The input ({\color{red}red}) and output ({\color{blue}blue}) states are Gaussian states. The non-uniform motion of the cavity produces two-mode squeezing. The procedure to enhance the effect is illustrated in the magnified {\color{Green}green} oval: by repeating $N$ times an arbitrary trajectory characterized by a total proper time $T$ (represented by the single box in the smaller oval at the top), the degree of squeezing is linearly increased.}

4 — 1201.1835

\caption{Upper bound to the capacity vs. rate $R$. Thresholds $G^*$ are reported for selected distributions. Distributions $\Lambda_i(x), i=1,\ldots, 5$ (\textcolor{red}{$\ast$}) are based on repetition codes ($k=1$). Distribution $\Lambda_6(x)$ ({$\square$}) is based on MDS codes ($k=2$). Distributions $\Lambda_i(x), i=7,8,9$ (\textcolor{magenta}{$\vartriangle$}) are based on MDS codes ($k=3$). Distributions $\Lambda_i(x), i=10,11$ (\textcolor{green}{$+$}) are based on MDS codes ($k=4$). Distributions based on $(k+1,k)$ SPC codes are also displayed (\textcolor{blue}{$\circ$}).}

5 — 1201.2410

\caption{Hess plot of disk sizes $(R_{\mathrm{disk}})$ of the late-type galaxies as a function of stellar masses $M_{\star}$. All galaxies shown are in the redshift range \redshift\and have not experienced a merger within the last 500 Myr. The right-hand panel shows only late-type galaxies that are brighter than 5 mJy in the SPIRE 250\mum band. The red lines show the median and $16$ and $84$ percentiles for simulated galaxies. The observed late-type galaxies, shown with green squares, are from \protect\cite{Cava:2010p1050}.}

6 — 1201.2644

\caption{\sl The allowed parameter space in the $(m_A,m_{H^\pm})$ plane (left panel) and $(m_\mu, \lambda_2)$ plane (right panel) taking into account theoretical and experimental constraints. The red dots (\redbullet) represent $R_{\gamma\gamma} < 1$ and blue dots (\bluebullet) represent $R_{\gamma\gamma} > 1$.}

7 — 1201.3050

\caption{(Color online) (a) Unit cell volume per formula unit (error bars are smaller than symbol size) dependence on $x$ of \lacorh\($\diamond$). The straight dashed line is a linear interpolation between $x=0$ and 1. (b) Average Co,Rh-O bond lengths ($\square$) and Co,Rh-O-Co,Rh angles (\textcolor{red}{$\circ$}) dependence on $x$. Corresponding values determined in Ref.~\cite{RefLi2010JSSC183_1388} are shown for comparison ($\times$).}

\caption{(Color online) Stabilization energy of LS/HS (full symbols) and IS (open symbols) configurations relative to LS configuration of \lacorh\for$x=0$ ($\square$) and $x=1/16$ (\textcolor{red}{$\circ$}) in dependence on cell volume by GGA+U calculation.}

\caption{(Color online) Stabilization energy of LS/HS relative to LS configuration of \lacotm\(M~=~Co, Rh, Al, Ga and In) for $x=1/16$ in dependence on ionic radius $r_M$ (the average for Co$^{3+}$ in LS and HS state is used for $r_{Co}$), by GGA+U calculation for room-temperature cell volume $V_{RT} $($\square$), $V/V_{RT}=99\%$ (\textcolor{red}{$\circ$}) and $V/V_{RT}=98\%$ (\textcolor{green}{$\triangle$}).}

8 — 1201.3263

\caption{{\bf Normal pressure fluctuations.} Dependence of the pressure tensor fluctuations on position along the $z$-axis for NEQ simulations at $P=3.6$, $T=4.0$ for $\tau=3\times 10^4$ ($\bigcirc$), $\tau=7\times 10^4$ ($\square$), and $\tau=9\times 10^4$ ($\Diamond$). The black line shows the value of the pressure tensor fluctuations for an equilibrium simulation in the smectic phase. Inset shows the local pressure integrated over the SD volume $\overline{\mathcal{P}}_{zz}$ ($\fullline$). The NEQ value oscillates with time following the external field, and, interestingly, its average value is close to the value characteristic of an {\it equilibrium} isotropic phase (\dashedline), which is the thermodynamic equilibrium state at this $P$ and $T$. The equilibrium value characteristic of the smectic phase is much larger (\dasheddottedline). From this we conclude that the local pressure is not large enough to drive the formation of a SD.}

9 — 1201.4005

\caption{\label{linlin} (Color online). Experimental measurements of contact radius $r$ plotted as function of time $t$ for different substrate wettabilities. Results for three different equilibrium contact angles are plotted: clean glass ($\theta_{e} \approx 0^{\circ}$; \textcolor{black}{\FilledDiamondshape}), coated glass ($\theta_{e} = 65 ^{\circ}$; \textcolor{red}{\FilledTriangleUp}), teflon coated glass ($\theta_{e} = 115^{\circ}$; \textcolor{blue}{\FilledCircle}). The curves represent averaged data of repeated measurements (five or more for each $\theta_{eq}$) per substrate for drops with radius $R=0.5$ mm, showing the reproducibility of the experiments. }

\caption{\label{EXPresults} (Color online). Experimental results. Contact radius $r$ measured as a function of time (top and right axis) for three different equilibrium contact angles: clean glass ($\theta_{e} \approx 0^{\circ}$; \textcolor{black}{\FilledDiamondshape}), coated glass ($\theta_{e} = 65^{\circ}$; \textcolor{red}{\FilledTriangleUp}), and teflon coated glass ($\theta_{e} = 115^{\circ}$; \textcolor{blue}{\FilledCircle}). The data shown, is an average of at least 5 measurements. The error bars denote the statistical error, which is larger than the measurement accuracy. On the left and bottom axis the data is normalized by the drop radius $R$ and inertial time $\tau_\rho = \sqrt{\rho R^3 /\gamma}$ respectively. A new regime is observed at earlier times, where the spreading is \textit{independent} of the equilibrium contact angle. The colored crosses are reprinted data by Bird et al.~\cite{Bird:2008} (corresponding with: $\theta_{eq} =$ $3^{\circ}$, \textcolor{black}{$\times$}; $43^{\circ}$, \textcolor{red}{$\times$}; and $117^{\circ}$, \textcolor{blue}{$\times$}). The arrow indicates the smallest times that were accurately resolved in the study by Bird~\emph{et al.}}

10 — 1201.4097

\caption{Measurements of (a)~optical depth and (b)~memory efficency of two \NdYSO{} crystals without~(\protect \bluecircle) and with~(\protect \redsquare) compensation scheme. Incident light is linearly polarized, and the $x$-axis indicates the angle of the polarization with respect to the $D_1$ axis of the crystals. (a)~Without compensation the optical depth varies between 2.70(1) and 0.99(1), corresponding to propagation along the two optical extinction axes. With compensation the peak-to-peak variation is reduced to 16\% of the mean optical depth. Lines are fits to Eq.~\eqref{eq:odepth}. (b)~Efficiency of 50~ns storage measured using laser pulses. With compensation the efficiency is almost independent of polarization.}

11 — 1201.4330

\caption{\label{fig:jobcenter} Dynamical load balancing in \reduze}

12 — 1201.5004

\caption{Azimuthal correlation at $|\deta|>0.7$ versus trigger azimuth relative to RP, $|\phis|$, in 20-60\% Au+Au collisions\red{~\cite{corrRP}}. The trigger and associated $\pT$ ranges are $3<\pTt<4$~\gev\and$1<\pTa<2$~\gev. The shaded areas are systematic uncertainties due to flow. The curves are Gaussian fit result: back-to-back ridges at $\dphi=0$ and $\pi$, and away-side peaks symmetric about $\dphi=\pi$.}

\caption{Away-side double-peak fit position (relative to $\pi$) as a function of (a) $\phis$ and (b) $\pTa$. Error bars are statistical only. \red{The systematic uncertainties due to elliptic flow are indicated by the dashed lines.} Data are from 20-60\% Au+Au collisions, and the trigger particle $\pTt$ range is $3<\pTt<4$~\gev\red{~\cite{corrRP}}.}

13 — 1201.5095

\caption{ \label{visc_vel_temp} (Color online) The logarithm of the effective viscosity as a function of the temperature for (a) ${\rm C}_{20}{\rm H}_{42}$, (b) ${\rm C}_{100}{\rm H}_{202}$ and (c) ${\rm C}_{1400}{\rm H}_{2802}$ system at four sliding velocities: 0.3 m/s ($\blacksquare$), 3 m/s ($\blacktriangle$), 10 m/s ($\blacklozenge$) and 100 m/s (\pentagon). }

14 — 1201.5448

\caption{\label{Fig:IndStocks:Level:5} Schematic illustration of coefficient significance in the power-law model (\ref{Eq:Pi:IP:Model:PL}) with $L=5$ for individual stocks: (a) buyer-initiated partially filled trades, (b) seller-initiated partially filled trades, (c) buyer-initiated filled trades, (d) seller-initiated filled trades. A plus or minus means that the corresponding coefficient is positive or negative, respectively. Red symbols ({\color{red}{$+$}} and {\color{red}{$-$}}) indicate that the corresponding variables are significant at the 5\% level, while black symbols ({\color{black}{$+$}} and {\color{black}{$-$}}) mean that the corresponding variables are insignificant at the 5\% level.}

\caption{\label{Fig:IndStocks:Level:2} Schematic illustration of coefficient significance in the power-law model (\ref{Eq:Pi:IP:Model:PL}) with $L=2$ for individual stocks: (a) buyer-initiated partially filled trades, (b) seller-initiated partially filled trades, (c) buyer-initiated filled trades, (d) seller-initiated filled trades. A plus or minus means that the corresponding coefficient is positive or negative, respectively. Red symbols ({\color{red}{$+$}} and {\color{red}{$-$}}) indicate that the corresponding variables are significant at the 5\% level, while black symbols ({\color{black}{$+$}} and {\color{black}{$-$}}) mean that the corresponding variables are insignificant at the 5\% level.}

15 — 1201.5562

\caption{\label{fig:temperature}(Color online) {\bf (a)} Fluctuations in the kinetic energy density, or the ``granular temperature'' $T_{\rm G}$, divided by the impact energy density scale $\rho_{\rm jet} U_0^2$ for half of the jet near the target. The granular temperature $T_{\rm G}$ is about $10^{-3}$ of the impact energy when $D_{\rm tar}/D_{\rm grain} = 200$. {\bf (b)} Normalized average granular temperature in the warm region above the target (\textcolor{blue}{\ding{58}}) and in the hottest region at the target edge (\textcolor{red}{\ding{54}}) plotted against the relative grain size $D_{\rm grain}/D_{\rm tar}$. By decreasing the grain size $D_{\rm grain}/D_{\rm tar}$, we suppress fluctuations within the jet.}

16 — 1201.5662

\caption{(\emph{Top}) Shift in the \tbroad-\thbr\plane as wavdetect's ellsigma parameter ranges from 0 to 3. Clusters from the B04 sample are depicted with solid arrows, while clusters from the D09 sample are depicted with long-dashed arrows. broadband spectral-fit temperatures are estimated within an aperture set at$R_{2500}$. Notice that the distribution of most of the points shifts significantly to higher spectral fit temperatures and smaller temperature ratios as cool lumps are excised. Notice also that some simulated clusters are unafflicted by cool lumps, so that their net shift is 0. (\emph{Bottom}) The \thbr\component (y-axis) of the shift presented in the top panel. Note that these are for simulated\xmas2 observations of maximum quality, having approximately 120K counts. \ColorRef}

17 — 1201.6251

\caption{Next\_chord\_prediction}

18 — 1202.0566

\caption{\label{fig:mu-diag}(Color online) The energy diagram. The \textcolor{blue}{blue} circles (\bluecircle) correspond to single-particle states, the \textcolor{red}{red} squares (\redsquare) to two-particle states and the \textcolor{brown!60!black}{brown} diamonds (\browndiamond) to three-particle states. The \textcolor{green}{green} solid line (\greensolid) is the bias window $\Delta\mu = \mu_L - \mu_R = 3.20-2.98 = 0.22$. The bias window includes the three particle ground state, but also excited one and two-particle states. So the expected number of electrons in the steady state is slightly below three.}

\caption{\label{fig:chg-cur-sin}(Color online) Charge and current for $\mu_L = 3.37$, $\mu_R = 3.15$, $u_C = 1.0$ and $\chi_\ell(t) \propto \sin(\omega t)$. Total charge \textcolor{brown}{brown} solid line (\brownsolid), charge for two particle states \textcolor{black}{black} dashed (\blackdash), for three particle states \textcolor{violet!60!black}{violet} dotted (\violetdot). Current for the left lead \textcolor{blue}{blue} dashed (\bluedash), for the right lead \textcolor{red}{red} solid (\redsolid). We consider two locations of the right lead. \textbf{(a)} Charge, left lead 1, right lead 10. \textbf{(b)} Current, left lead 1, right lead 10. \textbf{(c)} Charge, left lead 1, right lead 3. \textbf{(d)} Current, left lead 1, right lead 3.}

\caption{\label{fig:chg-cur-pulse}(Color online) Charge and current for the pulses. Total charge \textcolor{brown}{brown} solid line (\brownsolid), Charge for two particle states \textcolor{black}{black} dashed (\blackdash), Charge for three particle states \textcolor{violet!60!black}{violet} dotted (\violetdot), Current for the left lead \textcolor{blue}{blue} dashed (\bluedash), Current for the right lead \textcolor{red}{red} solid (\redsolid). $\mu_L = 3.37$, $\mu_R = 3.15$, $u_C = 1.0$, $\chi_\ell(t)$ pulses. \textbf{(a)} Charge left lead 1, right lead 10. \textbf{(b)} Current left lead 1, right lead 10. \textbf{(c)} Charge left lead 1, right lead 7. \textbf{(d)} Current left lead 1, right lead 7. \textbf{(e)} Charge left lead 1, right lead 3. \textbf{(f)} Current left lead 1, right lead 3. \textbf{(g)} Charge left lead 1, right lead 2. \textbf{(h)} Current left lead 1, right lead 2.}

\caption{\label{transf_charge}(Color online) Average current (Charge transferred) vs placement of the right lead. ${\mu_L = 3.37}$, ${\mu_R = 3.15}$, ${u_C = 1.0}$. The \textcolor{blue}{blue} dashed (\bluedash) line for $\chi_\ell(t)$ as rectangular pulses and \textcolor{red}{red} solid (\redsolid) for $\chi_\ell(t) \propto \sin(\omega t)$.}

19 — 1202.1821

\caption{{\color{red}{(Color online.) }} Examples of relaxed structures of the trans-PDT molecule bonded to gold clusters in different ways.\cite{Macmolplt} (a): Both S atoms bond to the gold clusters in the top geometry. (b): One S atom bonds to a gold cluster in the top geometry while the other S bonds in the bridge geometry. (c): Both S atoms bond to the gold clusters in the bridge geometry. (d): Both S atoms bond to the gold clusters in the hollow geometry. Carbon, hydrogen, sulfur, and gold atoms are black, grey, yellow, and amber, respectively.}

\caption{{\color{red}{(Color online.) }} Calculated IETS intensities and phonon energies for Mode-I for one or two PDT molecules connecting gold clusters together with the calculated low bias conductances. T: top site bonding, B: bridge site bonding, and H: hollow site bonding of a molecule to a gold clusters. TT and BB indicate two molecules connecting the gold clusters in parallel and bonding in top and bridge geometries, respectively. Gauche1: One gauche C-C bond. Gauche2: Two gauche C-C bonds. Numbers in the configuration column indicate the numbers of atoms in the gold clusters. Notice that the low bias conductance decreases with increasing numbers of gauche bonds for similar bonding geometries, consistent with previous theories for longer molecular wires.\cite{Paulsson.C.F.B.2009,LiPWTBAE2008} }

\caption{{\color{red}{(Color online.) }} \subref{Fig2(a)} Calculated vibrational modes in the phonon energy range from $38$ to $53$~meV for {\em trans}-PDT bridging gold nano-clusters with sulfur atoms bonded to gold in top-site and bridge-site geometries. Carbon, hydrogen, sulfur, and gold atoms are black, grey, yellow, and amber, respectively.\cite{Macmolplt} Red arrows show un-normalized atomic displacements; the heavier arrows indicate the motion of the sulfur atoms. Mode-I has the stronger IETS intensities. %%% \subref{Fig2(b)} Calculated IETS intensities (colored) vs. calculated phonon energies for {\em trans}-PDT molecules linking pairs of gold clusters with between 11 and 14 Au atoms in each cluster. Results are shown for both sulfur atoms bonding to the gold in top-site and bridge-site geometries and for top site bonding to one gold cluster and bridge site bonding to the other (bridge/top). The three ellipses enclose the calculated type I Mode-IETS spectra for pure bridge, pure top, and bridge/top bonding geometries for extended molecules with gold clusters containing various numbers of gold atoms. The experimental IETS phonon mode histogram of Hihath \emph{et al.}\cite{HihathArroyoRubio-BollingerTao08} is shown in (darker, lighter) grey for (positive, negative) bias voltages. Modes of the types I and II in (a) are indicated by arrows. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% }

\caption{{\color{red}{(Color online.) }} Calculated IETS intensities (colored) vs. calculated phonon energies for {\em trans}-PDT molecules linking pairs of gold clusters with 12 (yellow) and 13 (green) Au atoms in each cluster for pure hollow site bonding and with 13 Au atoms in each cluster for mixed hollow/top site bonding (magenta). The experimental IETS phonon mode histogram of Hihath \emph{et al.}\cite{HihathArroyoRubio-BollingerTao08} is shown in (darker, lighter) grey for (positive, negative) bias voltages. }

\caption{{\color{red}{(Color online.) }} \subref{Fig4(a)} Representative examples of calculated sulfur-related vibrational Mode-I and II of gauche-PDT molecules connecting gold nanoclusters each with {\color{red} 14} gold atoms, and the vibrational mode phonon energies. G1 and G2 label molecules with one and two gauche bonds respectively. The molecules connect to both gold clusters in pure bridge and pure top site bonding configurations. Carbon, hydrogen, sulfur, and gold atoms are black, grey, yellow, and amber, respectively. Red arrows show un-normalized atomic displacements. %% \subref{Fig4(b)} IETS intensities of gauche-PDT molecules linking pairs of gold clusters ({\color{red} 14} Au atoms in each cluster) vs. the vibrational mode phonon energy. Sulfur atoms bond to the gold in top-top, bridge-bridge site geometries. % The experimental IETS phonon mode histogram of Hihath \emph{et al.} \cite{HihathArroyoRubio-BollingerTao08} is shown in (darker, lighter) grey for (positive, negative) bias voltages. Modes of the types I and II are indicated by arrows. * Scaled down by a factor of 10. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% }

\caption{{\color{red}{(Color online.) }} The relaxed geometries and calculated conductances for pairs of trans PDT molecules connecting gold clusters in parallel. 14 gold atoms per cluster. (a)-I Top site bonding. (a)-II Bridge site bonding. Calculated mode frequencies and IETS intensities are shown in (b) for the systems depicted in (a). The calculated mode frequencies and IETS intensities for {\em single} molecules similar to those in (a) connecting two gold clusters with the same kinds of molecule-gold bonding geometries are also shown for comparison. The experimental phonon mode histogram of Hihath {\em et al.} \cite{HihathArroyoRubio-BollingerTao08} for {\em single} molecules bridging the electrodes is also shown. The darker (lighter) grey represents the positive (negative) bias voltage. }

20 — 1202.2465

\caption{The nested structure in the high school network represented as a Treemap. The color represents the best explaining attribute: \textcolor{blue}{\textit{blue}} for \textit{grade}; \textcolor{green}{\textit{green}} for \textit{race}; and \textcolor{yellow}{\textit{yellow}} for \textit{sex}. Numbers in parenthesis are the matching scores defined in the text. The size of shapes is proportional to the community size. Due to the page limit, only part of the entire treemap is shown.}

21 — 1202.2638

\caption{Comparison between math and CS\label{tb:mathvscs}}{\small \rowcolors{1}{}{LightSkyBlue} \begin{tabular}{l|l|l} Trends & CS & Math \\ \hline submitted more than one files & 84\% & 34\% \\ papers with no comments & 4.7\% & 9.6\% \\ average number of words in comments & 772 & 395 \\ number of pages most papers have & 6 & 10 \\ papers without any packages & 13\% & 25\% \\ average packages included in one paper & 6.7 & 5 \\ papers using \textsf{\textbackslash newcommand} &64\% & 66\% \\ average \textsf{\textbackslash newcommand} usages per paper &39.7 &36.1 \\ papers having theorems & 48\% & 71\% \\ \end{tabular} }

22 — 1202.5243

\caption{\coloronline (a) Schematic and (b) photograph of apparatus. The piston provides constant pressure via pulleys and weights, or constant volume by fixing its position to the surface of the table. (c) Entanglement of space-time trajectories for $N=20$ particles.}

23 — 1202.5250

\caption{Interferometer Performance (a) Measured phase (\protect \tikz{\protect \filldraw[fill=black!50, draw=black] (0,0) circle (2pt);}) and optical density (\protect \tikz{\protect \filldraw[draw=black!50,fill=white] (0,0) circle (2pt);}) as a function of detuning. Maximum saturation of the probe beam $s=1.2$, duration of the probe pulses $1.2\, \mathrm{\mu s}$. Solid lines result from a fit to the phase data using the model described in the text yielding $n_2\sigma_0=3.2$, effective saturation $0.6$ and an effective linewidth that is 20\% broader than the natural linewidth of $5.9\,\mathrm{MHz}$, which we attribute to the probe laser. Errorbars show standard deviations. (b) Measured phase-variance (without atoms) $\delta \phi^2$ (\protect \tikz{\protect \filldraw[fill=blue!50, draw=blue] (0,0) circle (2pt);}) as a function of photon number in the probe beam determined from 100 measurements for each point. The empty circle (\protect \tikz{\protect \filldraw[fill=white, draw=red] (0,0) circle (2pt);}) indicates the phase-variance for the intensity at which the spin-polarization measurements were made. Errorbars are an uncorrelated sum of statistical and systematic uncertainties. Expected phase-noise (\protect \tikz{\protect \fill[red!50!yellow!50] (0pt,2pt) -- (10pt,2pt) -- (10pt, 7pt) -- (0pt,7pt) --cycle;}) for quantum efficiency $\eta=0.6$. The width of the line corresponds to 20\% errors estimated from the uncertainty of our determination of $\eta$. The gray lines indicate the square of the phase shift expected for a single atom fixed in space at a detuning of half the atomic linewidth for the indicated $1/e^2$-waists of the probe beam \cite{aljunid_phase_2009}. }

\caption{Spin-Fluctuations in Ultracold Fermi Gases (a) Normalized histograms for the measured spin-polarization for weakly interacting gases at $T_1=1.02(2)\,T_F$ (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[blue!40!white,very thick] (0pt,2pt) -- (10pt,2pt);}), $T_2=0.44(5)\,T_F$ (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[blue!75!white,very thick] (0pt,2pt) -- (10pt,2pt);}) and $T_3=0.18(2)\,T_F$ (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[blue,very thick] (0pt,2pt) -- (10pt,2pt);}) as well as for a strongly interacting gas of molecules at $T_{{\rm mol}}=0.16(7)\,T_F$ (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[red,very thick] (0pt,2pt) -- (10pt,2pt);}), where $T_F$ is the Fermi temperature. (b) Reduction of spin-fluctuations compared to a thermal gas of equal column density as a function of $T/T_F$. Data points are colored identical to (a). The solid line shows theory for a noninteracting Fermi gas. (c) Spin susceptibility per atom $\chi_{\rm col}/n_{\rm col}$ as a function of temperature for the weakly interacting gas (blue points) and the strongly interacting gas (\protect \tikz{\protect \filldraw[fill=white, draw=red] (0,0) circle (2pt);}). Also shown are the susceptibility for a classical gas (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[black!50!white,very thick, dashed] (0pt,2pt) -- (10pt,2pt);}), which varies as $1/T$ and the theoretical expectation (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[blue!40!white,very thick] (0pt,2pt) -- (10pt,2pt);}) for the trap parameters of the gas at $T_3$. (d) Spin-fluctuation density $A \delta m^2=kT\chi_{\rm col}$ as a function of column-density $n_{\rm col}$. The prediction for a thermal gas (\protect \tikz{\protect \draw[white] (0pt,0pt) -- (10pt,0pt); \protect \draw[black!50!white,very thick, dashed] (0pt,2pt) -- (10pt,2pt);}) is shown along with the boundary separating the separable from the non-separable region. The blue line shows theory for a noninteracting Fermi gas at temperature $T_{{\rm mol}}=0.36\,{\rm \mu K}$. Errorbars are an uncorrelated sum of statistical and systematic uncertainties. }

24 — 1202.5675

\caption{\redtw(graph $H$, required vertices $R$, boundary vertices $B$)}

25 — 1202.5687

\caption{{\bf Experimental setup and protocol.} (color online) \leg{(a):} The vibrating cell. \leg{(b):} Logarithmic increase followed by a stepwise decrease of the packing fraction, and pressure at the wall. \leg{(c)}: Average pressure vs packing fraction, $\gamma=1.4$: (\textcolor{blue}{$\bigcirc$}) : $P_{TOT}$, (\textcolor{red}{$\square$}): $P_{STAT}$, (\textcolor{green}{$\triangle$}): $P_{DYN}$ as defined in the text. }

26 — 1202.6063

\caption{Semimajor axis (top panel in each plot) and eccentricity (bottom panel in each plot) evolution for all the four runs. In each panel, the \textcolor{red}{red} line represents the full evolution directly taken from the simulation whereas the \textcolor{black}{black} line is the equivalent evolution when considering the energy and angular momentum exchanges due to accreted particles only (i.e., basically due the evolution of the mass and mass ratio of the binary).}

\caption{Consistency check of the evolution of the binary angular momentum components for all the simulations. In each plot, the $L_x$, $L_y$ and $L_z$ component evolution is shown from the top to the bottom. In each panel, the dashed line is the $\Delta L$ exchange balancing the gravity torques at each time step, the dotted line is the $\Delta L$ exchange computed from the accreted particles, and the solid black line is the sum of those two. For comparison, the \textcolor{red}{red} line is the angular momentum evolution taken directly from the raw simulation data. All $\Delta L$ are in units of $[M_0a_0V_{c,0}]$.}

\caption{Decomposition of the $L_z$ evolution for the four runs. The $\Delta{L_z}$ contributions stem from \textcolor{blue}{$a$ (dotted)}, \textcolor{green}{$e$ (dashed)}, \textcolor{Plum}{$M$ (dot dashed)} and \textcolor{BurntOrange}{$\mu$ (long dashed)} variations separately. The sum of the four contributions is given by the solid black line, whereas the solid \textcolor{red}{red} line is, as in Fig. \ref{lz_evol}, the angular momentum evolution taken directly from the raw simulation data. All $\Delta L$ are in units of $[M_0a_0V_{c,0}]$.}

\caption{Differential torques $dT/dr$ in units of $[G M^2_0a^{-2}_0]$ averaged over the entire simulations. In each panel we show the differential torque on the primary (\textcolor{green}{green}), on the secondary (\textcolor{red}{red}), the sum of the two (\textcolor{blue}{blue}), and the total integrated torque (\textcolor{black}{black}) according to Eq. \ref{eqavtorque}. This latter is integrated starting from $a$ inwards and outwards. Notably, the inward torque is positive, whereas the outward torque is negative. }

\caption{Disc properties: average surface density (top panel), and Toomre parameter $Q$ (bottom panel) as a function of $r$. Line style denotes \textcolor{red}{iso10} (long dashed), \textcolor{red}{iso05} (dot dashed), \textcolor{black}{adia10} (solid) and \textcolor{black}{adia05} (dashed). }

\caption{Simple test model for the periodicity generated in the outside disc, i.e. at $r>a$. In the top panel we show the temporal evolution of the torque, whereas in the bottom panel we show the Fourier transform. In each panel, the \textcolor{blue}{blue} curve is obtained by placing three point masses at distance $1.6a$, $2.1a$, $3.5a$ from a circular binary, and the one \textcolor{red}{red} is the torque found in the {\it iso10} simulation. }

\caption{Accretion onto the two BHs. Top box: accretion rate as a function of time. The \textcolor{red}{red} line is the total accretion rate while the \textcolor{blue}{blue} and \textcolor{black}{black} lines are the accretion rates onto the secondary and the primary respectively. Lower box: Fourier transform of the total accretion rate (lower panel) and of the torques (upper panel). In the upper panel, the \textcolor{green}{green} line is the total torque (highest peak normalized to $1$) and the \textcolor{red}{red} line is the torque exerted by the material at $r<a$ (multiplied by ten). }

27 — 1202.6569

\caption{Characterization of the ejecta regimes. (\textcolor{blue}{$\CIRCLE$}) Smooth ejecta sheet. ($\blacksquare$) Irregular splashing. ($\mathrlap{\textcolor{yellow}{\blacktriangledown}}\triangledown$) Bumping. (\textcolor{red}{$\blacktriangle$}) Quartering. (\textcolor{green}{$\blacklozenge$}) Protrusions rising up along the side of the drop.}

28 — 1203.0386

\caption{Distribution of the arrival directions of UHECR, represented by black dots ($\bullet$), with energy $E\geq55\,{\rm EeV}$ reported by PAO in 2010, in the equatorial coordinates plotted using the Hammer projection. The solid red line represents the boundaries of the sky covered by PAO experiments. The blue crosses ({\color{blue}$\times$}) represent the locations of AGN with distance $d \leq 100\,{\rm Mpc}$ taken from the 13th edition of VCV catalog. The cyan square ({\color{cyan}$\blacksquare$}) and the cyan triangle ({\color{cyan}$\blacktriangle$}) show the locations of Centaurus A and Messier 87, respectively.}

29 — 1203.0672

\caption{(Color online) Calculated and experimental pressure-volume relations for solid H$_2$ and D$_2$. Semi-empirical calculations for many-body potential (this work): $p-$H$_2$ (solid line), $o-$D$_2$ (dashed line); DFT-GGA calculations ($\cdot \cdot \cdot$) \cite{Zhan07}. Experiment (reduced to 0 K): (H$_2$: {\color{green}$\blacktriangle$}\cite{Aka10}, {\color{red} $\circ$} \cite{Lou96}, {\color{green} $\square$} \cite{Mao94}); (D$_2$: {\color{blue} $+$} \cite{Lou96}, {\color{blue} $\times$} \cite{Bes90}, {\color{blue} $\ast$} \cite{Mao94}; the inset shows the small-pressure range. Experiment (reduced to 0 K): (H$_2$: {\color{red} $\circ$} \cite{Dri79}, {\color{red} $\times$} \cite{vS88}, {\color{red} $\triangledown$} \cite{Ish83}); (D$_2$: {\color{blue} $\square$} \cite{Dri79}, {\color{blue} $+$} \cite{vS88}, {\color{blue} $\ast$} \cite{Ish85}.}

\caption{(Color online) Calculated and experimental Raman frequencies as a function of pressure for solid hydrogen and deuterium. Semi-empirical calculations for many-body potential: this work. $p-$H$_2$ (dashed line), calculations for the SG potential {\color{MyOrange} $\ast$} \cite{Mao94}; $o-$D$_2$ (solid line). DFT-LDA theory (this work): H$_2$ $\circ$, D$_2$ $\bullet$. Experiment (H$_2$): {\color{blue} $\diamond$} \cite{Han94}, {\color{red} $\times$} \cite{Hem90a}), {\color{green} $\square$} \cite{Gon98,Gon01}; (D$_2$): {\color{red} $\triangledown$} \cite{Hem93}.}

30 — 1203.1626

\caption{ %{\bf AN ALTERNATIVE VERSION OF Figure1 SHOWING THE FREE-FREE CORRECTED RADIO MAP} The point-source subtracted 1.4~GHz (left) and 24~$\mu$m (right) maps surrounding 30~Dor with residual $1-3$~GeV \gray{} contours starting at $0.5 \times 10^6$ counts sr$^{-1}$ (roughly $\sim$1 count per $0.1\degr \times 0.1\degr$ pixel) with linear intervals of $0.326 \times 10^6$ counts sr$^{-1}$ overlaid. The 1.4~GHz map has been corrected for free-free emission. The region shown is centered at RA~$05^{\rm h}~39^{\rm m}~07^{\rm s}$, Dec.$-69^\circ 22\arcmin 02\arcsec$ (J2000) and has a radius of 1$^\circ$ ($\approx875$~pc). The residuals between the radio/\gray{} and infrared maps were calculated over this region ($\S$\ref{sec-imsm}), as well the minimum energy magnetic field ($\S$\ref{sec-datradcre}), and the radiation field energy density ($\S$\ref{sec-imsm}). In addition, we show the locations of two Crab-like pulsars, PSR~J0540$-$6919 (square) and PSR~J0537$-$6910 (circle), along with a known background X-ray source RX~J0536.9$-$6913 (`X') that are coincident with the line-of-sight towards 30~Dor on the 1.4~GHz and 24~$\mu$m maps. \label{fig-1}}

\caption{ The same region on the sky as shown in Figure~\ref{fig-1}, but now displaying the H{\sc i} column density map \citep{lss03,sk03}. %Residual $1-3$~GeV (left) and $3-10$~GeV (right) \gray{} contours are overlaid with the same intervals as shown in Figure~\ref{fig-1}, and In the left panel, residual $1-3$~GeV \gray{} contours are overlaid on the H{\sc i} column density map with the same intervals as shown in Figure~\ref{fig-1}. In the right panel, residual $3-10$~GeV \gray{} contours are overlaid on the H{\sc i} column density map starting at $0.2\times10^{6}$~counts~sr$^{-1}$ with linear intervals of $0.178\times10^{-6}$~counts~sr$^{-1}$. Also shown are the two Crab-like pulsars, PSR~J0540$-$6919 (square) and PSR~J0537$-$6910 (circle), along with a known background X-ray source RX~J0536.9$-$6913 (`X') that are coincident with the line-of-sight towards 30~Dor. \label{fig-2}}

\caption{ The residuals between the free-free corrected 1.4~GHz and smoothed 24~$\mu$m (as defined in \S \ref{sec-imsm}) plotted as a function of both exponential (asterisks) and Gaussian (diamonds) of kernel scale lengths. The shaded region indicates the extrapolated propagation length for GeV protons based on the compact morphology of the \gray{} emission observed by the \fermilat\\citep{aa10-lmc}. The value reported by \citet{aa10-lmc} is the $\sigma$ from a modeled Gaussian. %It has therefore been multiplied by $\sqrt{2}$ for proper comparison with the best-fit scale length from the Gaussian kernel used in the present analysis. The values for the Gaussian scale lengths agree within errors. \label{fig-3}}

\caption{ The residuals between the $1-3$~GeV \gray{} and smoothed 24~$\mu$m images (as defined in \S \ref{sec-imsm}) plotted as a function of both exponential (asterisks) and Gaussian (diamonds) of kernel scale lengths. The asterisk and diamond attached to the horizontal dotted lines indicate the location of the best-fit scale length measured for the CR electrons using exponential and Gaussian kernels (see Figure~\ref{fig-3}). \label{fig-4}}

31 — 1203.2581

\caption{(Color online) (a) The schematic crystal structure of KFe$_2$As$_2$; (b) Sketch of the lattice structure of FeAs layers. The As ions {\textcolor{green}\textbullet} and {\textcolor{magenta}\textbullet} are located just above and below the center of each face of the Fe square lattice, respectively.}

32 — 1203.2718

\caption[]{\label{fig:stills} a) Images of a $10$ cSt silicone oil drop with $u_{0}$=$4.1$ m/s (Re=$1300$, We=$2400$) impacting acrylic surfaces with root-mean-square roughness $0.02$ $\mu m$ (left), $1.05$ $\mu m$ (middle), and $2.84$ $\mu m$ (right) at atmospheric pressure. The images were taken $0.25$ ms (top), $0.75$ ms (middle) and $1.50$ ms (bottom) after impact. The smoothest surface produces a thin-sheet splash, the roughest produces a prompt splash, while the intermediate roughness surface completely suppresses splash formation. b) Impact of a $5$ cSt silicone oil drop (Re=$2500$, We=$2400$) on acrylic with $R_{rms}$=$1.05$ $\mu m$. A prompt splash is formed, followed by thin-sheet ejection and splash. c) An ethanol drop ($\nu$=$1.36$ cSt) at $u_{0}$=$3.1$ m/s (Re=$7100$, We=$1050$) shown $0.25$ ms after impact on rough acrylic ($R_{rms}$=$1.05$ $\mu m$): at $P$=$101$ kPa (left), $49$ kPa (middle) and $15$ kPa (right). Prompt splashing is suppressed at lower gas pressure. All images at $t=0.25$ ms are magnified to show prompt splashing. d) Roughness power spectrum, $H(k)$, for rough acrylic with $R_{rms}$=$0.34$ $\mu m$ (\includegraphics[height=6pt]{RedCircle.pdf}), $R_{rms}$=$1.05$ $\mu m$ (\includegraphics[height=6pt]{BlueTriangle.pdf}) and $R_{rms}$=$2.84$ $\mu m$ (\includegraphics[height=6pt]{MagentaPlus.pdf}) and of sandpaper with $R_{rms}$=$2.47$ $\mu m$ (\includegraphics[height=6pt]{BlackSquareFilled.pdf}). The surfaces have the same qualitative shape of $H(k)$. }

\caption[]{\label{fig:acrylic_phase_diagram}Splashing phase diagram for silicone oil drops impinging on an acrylic surface at $u_{0}$=$3.4$ m/s ($730$<Re<$1900$, We$\approx$1650) and atmospheric pressure. At low roughnesses, the impact results in a thin-sheet splash (\includegraphics[height=6pt]{BlueSquare.pdf}) and at high roughnesses - a prompt splash (\includegraphics[height=6pt]{RedTriangle.pdf}). At intermediate roughnesses two outcomes are possible. At lower viscosities, a prompt splash is followed by a thin-sheet splash in the same impact event (\includegraphics[height=6pt]{BlueSquareRedTriangle.pdf}). For higher viscosities, splashing is suppressed completely (\includegraphics[height=6pt]{BlackCircle.pdf}): one can suppress splashing by \emph{increasing} the roughness of the surface. The dashed lines separating different splashing regimes are guides to the eye, and shift in the direction given by the arrows as air pressure is reduced. }

\caption[]{\label{fig:ejection_times} Thin-sheet ejection times vs. surface roughness for silicone oil drops of viscosity $5.5$ cSt (\includegraphics[height=6pt]{BlackCircle.pdf}), $10.3$ cSt (\includegraphics[height=6pt]{BlackTriangle.pdf}), and $14.4$ cSt (\includegraphics[height=6pt]{BlackSquare.pdf}) impacting an acrylic surface at $u_{0}$=$3.4$ m/s ($730$<Re<$1900$, We$\approx$1650). As $R_{rms}$ increases, $t_{ejt}$ increases. For the higher viscosities and $R_{rms}$>$0.23$ $\mu m$, the thin sheet no longer breaks apart into smaller droplets (open symbols).}

33 — 1203.2861

\caption{Constructional details of the hand held digital IC tester probe in a empty container of glue stick tube(8-15g). The pulse switch(T) and the queiscent state switch(S) are at 90$^o$ to each other. R$_2$=680k$\Omega$ and R$_3$=220k$\Omega$ correspond to the logic probe[2]. They are in series with the R$_1$ and R$_4$ resistors. In other words their black and red wires connect to the pin-12 and pin-8 of HEF4011BP. The LEDs have 470$\Omega$ resistors in series. {\color{red}L-}/{\color{green}L+} are connected to GND/+5V respectively.}

34 — 1203.3221

\caption{\textbf{Left:} Effective temperature of rotational \HBM. The simulation results for $\THBM$ ({\large \color{myblue} $\bullet$}) were deduced from the numerically measured $\zetaHBM$ and $\DHBM$ using the generalized Einstein relation \eqref{eq:GER}. An alternative estimation of $\THBM$ ({\color{myblue}\footnotesize $\square$}) was obtained from the Boltzmann distribution of the inclination angle $\theta$ in a harmonic angular confinement (right panel). The theoretical prediction (solid line) was evaluated within the idealized theory for an incompressible fluid via \fref{eq:T_HBM_rot_final} using the radial viscosity profile $\eta(r)$ determined in the MD simulation. For comparison, the effective temperature $T_\mathrm{HBM}^\mathrm{t}$ for the translational degrees of freedom (dot-dashed line) \cite{Chakraborty:2011} and the solvent temperature at the particle surface (dotted line) are shown. \textbf{Right:} The measured distribution of the inclination angle $\theta$ in a harmonic angular confinement potential for nanoparticle temperatures $T_\mathrm{p} =0.8\,\epsilon/k_\mathrm{B}$ ({\color{myblue}{\large $\bullet$}}), $1.25\,\epsilon/k_\mathrm{B}$ ({\color{mypurple}\footnotesize $\blacksquare$}) and the corresponding distribution $p(\theta) \sim e^{-\beta V(\theta)}$ with $\beta^{-1}=k_\mathrm{B}\THBM$ depicted by the solid lines.}

\caption{Radial variation of the measured angular velocity $\w(r)/r$ normalized by the angular velocity of the hot nanoparticle for $T_\mathrm{p}=0.9 \epsilon/k_\mathrm{B}$ ({\color{myblue}{\large $\bullet$}}) and $1.25 \epsilon/k_\mathrm{B}$ ({\color{mypurple}\footnotesize $\blacksquare$}). The solid lines are the corresponding plots of \fref{eq:gensol_gen_problem} evaluated using the measured viscosity and temperature profiles. \emph{Inset:} The variation of the angular velocity near the nanoparticle surface. The vertical solid lines indicate the positions of the hydrodynamic boundary condition for the predicted hydrodynamic flow fields \eqref{eq:gensol_gen_problem} fitted to the simulation data. The shaded region (its extension marked by an arrow) indicates the equivalent sphere radius for the nanoparticle (dot-dashed line) plus $\sigma/2$. The comparison reveals a weak apparent slip at high nanoparticle temperatures due to the radial solvent density variation induced by the heating.}

35 — 1203.4208

\caption{ (left panel) A thumbnail representation of a WFI \reduc\created by STIFF. The optimized intensity cuts and binning allow a quick assessment of the quality. This particular example shows an intensity gradient caused by either poor flat fielding, nebulosity from a galaxy at the center of the mosaic field (to the upper left), or simply a non-uniform illumination of the focal plane. The intensity values are inverted. (right panel) A thumbnail representation of a WFI\weigh\created by STIFF. The optimized intensity cuts and binning allow a quick assessment of the quality. This particular example is associated with the thumbnail in the left panel. Saturated stellar peaks and bad columns are clearly visible in addition to ``doughnuts'' of the primary mirror of the telescope that are part of the flat field foundation of the\weigh. White pixels have values near 1, black pixels have values at or near 0. The horizontal lines are artifacts of the CCD manufacturing process. The higher weight of the pixels near some of the bad columns is an artifact caused by Fourier processing of input flat frames without properly taking into account bad pixels. It is possible to identify some of these defects with pixel statistics \textit{a priori}, but these unusual cases are generally only identified through this type of inspection plot. }

\caption{ A PSF Anisotropy plot of the \reduc\whose thumbnail is seen in Fig.~\ref{fig:ithum}. % Taking into account figure rotation, t The left panel shows the two-dimensional anisotropy in the PSF (in both FWHM and in ellipticity) of the sources. The top-middle panel shows source magnitudes on an arbitrary scale versus flux radius and gives an indication of how ``stellar'' the sample is. The remaining panels plot FWHM of the sources versus ellipticity position angle, horizontal position and vertical position, respectively, clockwise from top-right. }

\caption{ An \astro\inspection plot for a recent solution of the\reduc\associated with the thumbnail in Fig.~\ref{fig:ithum}. The plot displays the statistics of the residuals (DRA and DDEC) between the RA and DEC of sources in a source catalog to which the local astrometric solution has been applied and the RA and DEC of those sources as listed in the reference catalog of astrometric standards, USNO-A2.0 in this case. The text in the top of the figure lists the observation date (DATE\_OBS), the number (N) of sources pairs plotted, their average RA ($<$RA$>$) and DEC ($<$DEC$>$) in degrees, the average RA and DEC residuals ($<$DRA$>$ and $<$DDEC$>$) and their standard deviations in arcsec, and finally the root-mean-square (RMS) of the two-dimensional residual and the maximum two-dimensional residual (Max) in arcsec. The large upper panel plots DRA versus DDEC. The four panels below it show DRA and DDEC with respect to RA (with a constant offset of 203.9 degrees) and then to DEC. }

\caption{ A \photo\inspection plot for a photometric observation comprising one WFI detector. A graphical representation of the data used to calculate the photometric zeropoint. In this plot, three photometric reference catalogs can be seen: Stetson (blue points),\aw\secondary standards (red points) and Sloan Digital Sky Survey data release 5 (black points).}

\caption{ Plot of the dome flat exposure level (median value of the science region minus the median value of the X overscan region) versus exposure time of ESO\_CCD\_\#65 of the OmegaCAM instrument from data taken in 2011. This plot gives a quick indication of how linear this detector is. The dashed red line is only an indication of the slope in the data. The cluster of points at EXPTIME$=$3 sec is from heavier sampling for diagnostic and detector health purposes. }

\caption{ Plot of the dome flat exposure (median value of the science region minus the median value of the X overscan region divided by the exposure time) versus the dome flat exposure level (median value of the science region minus the median value of the X overscan region) of ESO\_CCD\_\#65 of the OmegaCAM instrument from data taken in 2011. This plot quickly gives a different view of how linear this detector is. The dashed red line is only an indication of the mean detector exposure.}

\caption{ A plot of delta MAG\_AUTO (blue points) over-plotted with delta DEC (red points) versus MAG\_AUTO. The increase in scatter of the astrometric residuals is far lower than that of the photometric residual, a qualitative indication that astrometry for faint sources is at acceptable levels.}

\caption{ Screen-shot of the upper part of a \qwise\page. This view shows the quality of an OmegaCAM coadded frame. At the very top is the type of object and a link to the file on the dataserver (a unique hash value in the filename link is purposely obscured for security reasons). Directly below the banner is the top bar with links and basic actions. Below this is tabular information about the object and graphical inspection plots (a thumbnail of the image on the left and its weights on the right, cf. Fig.~\ref{fig:ithum}). Note that green fields indicate values within specified ranges that will be red when out of specified ranges. }

36 — 1203.4296

\caption{\label{fig:switching_times} Switching times for orders $n=1$ to $n=10$. UDD switching times are open circles \opencircle. Even permutation $A_3$ DFS DD switching times are filled circles \fullcircle only. Switching times for $S_3$ DFS DD over all six $S_3$ permutations include open circles, filled circles, and stars $\star$. The switching times are symmetric about $t=\frac{1}{2}$.}

\caption{\label{fig:classical_bath} Simulations of a DFS qubit coupled to a classical dephasing-only bath. Infidelity is plotted against total evolution time $T$ for $A_3$ DFS DD orders $n=0$ (free evolution) through $n=4$. $n=0$ infidelities are given by \fullcircle, $n=1$ by \fullsquare, $n=2$ by $\blacklozenge$, $n=3$ by $\blacktriangle$, $n=4$ by $\blacktriangledown$. Dotted lines are fits to the given leading order infidelity term.}

\caption{\label{fig:quantum_bath} Simulations of a DFS qubit coupled to a quantum bath. Infidelity is plotted against total evolution time $T$ for DFS DD orders $n=0$ (free evolution) through $n=3$. $n=1$ and $n=2$ use $S_3$ DFS DD pulse sequences; $n=3$ uses the pulse sequence in table \ref{tab:qdd3}. $n=0$ infidelities are given by \fullcircle, $n=1$ by \fullsquare, $n=2$ by $\blacklozenge$, $n=3$ by $\blacktriangle$. Dotted lines are fits to the given leading order infidelity term.}

37 — 1203.5158

\caption{Across 1-year intervals, 1985--2007: {\sc(a)} Number of publications. {\sc(b)} Number of authors. {\sc(c)} Number of publications across subnetworks. {\sc(d)} Number of authors across subnetworks. {\sc(e)} Average number of authors per publication. {\sc(f)} Average number of subject classifications per publication. {\sc(g)} For each fixed number of authors (1--8), a histogram over years of publications attributed to that number of authors. Each histogram is scaled by the total number of publications by that number of authors. We include four subdisciplines: algebra ({\color{alg}\(\circ\)}; 08--22), differential equations ({\color{diffeqn}\(\scriptstyle\square\)}; 34--35), computer science and information ({\color{compsciinfo}\(\diamond\)}; 68, 94), and classical physics ({\color{phys}\(\scriptstyle\triangle\)}; 70--86).\label{win1}}

38 — 1203.5982

\caption{{\it Perpetual power-law coarsening for $j'_E>0$:} (a) The black line indicates the steady state amplitude $A$ versus wavelength $\lambda$ of periodic steady-states for $j'_E = 0.45$. Crosses ({\color{red}$\times$}): full simulations from small random initial conditions. (b) Rescaled Amplitude $A/j'_E$ and wavelength $\lambda/(j'_E)^{1/2}$ as function of time for starting from small random initial conditions. The solid lines are the powerlaws discussed in the text.}

39 — 1203.6240

\caption{The best-fit parameters of our geometric modeling. $R_{\mathrm{\highlight{ring}}}$ is the inner radius of the ring-shaped disk models. $f_{\mathrm{\highlight{ring}}}$, $f_{\mathrm{star}}$, and $f_{\mathrm{halo}}$ denote the flux contributions of the \highlight{ring}, the star, and the halo, respectively. $i$ is the inclination angle ($0\degree$ corresponds to a face-on \highlight{ring}). $\theta_{\mathrm D}$ is the position angle of the \highlight{ring}'s major axis. $\chi^2_{\mathrm{red}}$ is the reduced chi-square. }

40 — 1204.0059

\caption{(color online). Average global 4-fold bond-orientational order parameter $\langle |Q_4| \rangle$ versus $\Gamma$ for C1 (a) and C2 (b). Open (solid) symbols represent data obtained for increasing (decreasing) $\Gamma$ ramps, with the following rates: $\Delta\Gamma/\Delta t \approx 0.005$ min$^{-1}$ ({\color{blue} $\triangle$, $\blacktriangledown$}) and $\Delta\Gamma/\Delta t \approx 0.02$ min$^{-1}$ ({\color{red} $\circ$, $\bullet$}). Continuous lines in (b) correspond to fits of the linear trend $Q_4^L = a\Gamma +b$ for $2.5<\Gamma<5$, with $a=0.011\pm0.001$ and $b=0.380\pm0.002$, and a supercritical-like behavior $ \langle |Q_4| \rangle=Q_4^L + c(\Gamma-\Gamma_c)^\beta$, with $\beta=1/2$, observed for $\Gamma \gtrsim 5$. }

41 — 1204.0271

\caption{Fender's Blue Butterfly Habitat - Kincaid's Lupin Patches - U.S. Fish and Wildlife Service,Wikipedia http://en.wikipedia.org/wiki/Fender\%27s\_blue\_butterfly, http://www.fws.gov/oregonfwo/Species/PrairieSpecies/gallery.asp}

42 — 1204.4676

\caption{\label{fig: T1T2}Variation of the function $F(x_3)=({4\over 3})T_1T_2x_3^2M_2^2/\omega_L^2$ as a function of $^3$He concentration. For a unique correlation time $F(x_3)=1.0$. Experimental data: \fullsquare \, Kim{\it et al.} \cite{SSK-PRL2011}, {\textcolor{blue} \opentriangle \, Allen{\it et al.} \cite{Allen82}}, \textcolor{orange} \fullcircle \, Schratter{\it et al.} \cite{Schratter84},\,\opendiamond \, Schuster{\it et al.} \cite{Schuster}, \textcolor{green} \opentriangledown \, Hirayoshi{\it et al.} \cite{Hirayoshi}.}

\caption{\label{fig: CT1}Observed concentration dependence of the nuclear spin-lattice relaxation times for dilute $^3$He in solid $^4$He. \textcolor{red}\opensquare \, Greenberg{\it et al.} \cite{1972JLTP...8..3} and for the rest of symbols see the caption for Fig. \ref{fig: T1T2}.}

43 — 1204.5600

\caption{Comparison of predicted stellar parameters %\Teff--\logg\ relations for theoretical T6.09 isochrones and the stellar samples. Symbols represent the various theoretical temperature scales (see text). Shown as crosses are results with error bars from the spectroscopic analysis of \citetalias{korn2007}. Note that the \SI{13.5}{Gyr} model was computed at \SI{0.1}{dex} higher metallicity, which produces a slight deviation on the RGB. The difference is negligible near the TOP. \coloredition}

\caption{Chemical abundances on the shifted temperature scale \scaleshifted, compared to predictions from \SI{12.5}{Gyr} isochrones. Dashed horizontal lines represent the initial composition. The stellar models include the effects of atomic diffusion with a free parameter for the \tme, at five different values. We find that models in the range T5.95--6.15 all reproduce observations quite well. The optimal model is T6.0. \coloredition\label{fig:trends-100s}}

\caption{Observed abundance trends in lithium, shown for the shifted temperature scale \scaleshifted. Lines represent the predicted surface abundance evolution at four different efficiencies of turbulent mixing. The horizontal dashed line represents the original lithium abundance of all four models, the dotted line the predicted primordial abundance. The shaded areas represent the respective uncertainties, and their overlap. \coloredition\label{fig:lithium}}

44 — 1204.5972

\caption{(color online). The average orientational angle $\theta_{\rm av}$ as a function of the (a) length to diameter ratio $L/d$ of the particles and (b) order parameter of the system obtained by experiments ($\bullet$,$\blacktriangle$,$\vartriangle$) and numerical simulations ({\color{red}$\times$}). One datum point ({\color{blue}$\blacktriangle$}) was taken with a particularly deep layer, $H/d=34$. The alignment angles for typical flow aligning nematic liquid crystals MBBA and 5CB ({\color{darkgreen}\myeightstar,{\large$\diamond$},+}) \cite{skla1979,ga1972,beje1985,chch2004,ehhe1995} and for a nematic liquid crystal 8CB ({\color{blue} $\Box$}) which shows tumbling when approaching the smectic A phase transition are shown for comparison. }

45 — 1204.6051

\caption{Three-body losses for exactly two impurity atoms, observed event-by-event. \textbf{(a)} Fluorescence traces showing all three possible loss events and the according allowed loss channels. \textbf{(b)} Corresponding relative occurrences for all possible case: two atom loss (\textcolor{red}{$\blacktriangle$}), one atom loss (\textcolor{darkgreen}{$\blacktriangledown$}) and no atom loss (\textcolor{blue}{\textbullet}). The solid lines show the expectation, for details see text.}

46 — 1204.6248

\caption{ \textbf{An example of model dynamics starting from the initial configuration (7,5). In red, the pairing that it is established at each step.} (a) At time $\tau$, $I=7$ infected mobile phones $b_1,\ldots,b_7$ and $S=5$ clean mobiles $w_1,\ldots,w_5$ are all within their mutual Bluetooth connection range. (b) $b_1$ chooses a mobile among $b_2,\ldots,b_7,w_1,\ldots,w_5$; it chooses $w_1$ establishing connection \ding{202}. (c) Now it $b_2$'s turn to choose, and $b_1$ and $w_1$ are not available anymore for pairing (marked by a grey circle \textcolor{gray}{\CircleSolid}). (d) $b_2$ connects to $b_3$ through pairing \ding{203}. (e) The two mobiles $b_2$ and $b_3$ become unavailable for pairing, too and the next infected mobile in line $b_4$ pairs to $w_2$ via \ding{204}. (f) Only $b_6, b_7$ and $w_3,w_4,w_5$ remain available for pairing with $b_5$, which chooses $b_7$ (connection \ding{205}). (g) Now the last mobile $b_6$ must connect to the remaining unpaired clean phones $w_3,w_4,w_5$: it chooses $w_4$ creating pairing \ding{206}. (h) There are no more unpaired infected mobiles: the process ends at time $\tau+\Delta\tau$. }

47 — 1205.0238

\caption{(color online) (a)~Distribution of the apparent entropy production $p(\sred)$ for different trajectory lengths $t$ and plasma parameters $\Gamma$. (b)~Section of previous histograms around $\sred=0$. (c)~Corresponding $\ln\left[p(\sred)/p(-\sred)\right]$ as a function of $\sred$. The dashed black line has the theoretically predicted slope $1$, whereas the red line is a linear fit with slope $\alpha = 0.65$. \label{fig2}} \end{figure} The black histograms (closed bars and line) in Fig.~\ref{fig2}(a) show the distribution of the apparent entropy production $p(\sred)$ in the absence of coupling obtained for trajectories of length $t=1.75\unit{s}$ and $10\unit{s}$, respectively. The peaked distribution shifts with elapsing time to the right with peak height maxima occurring at positions which correspond to the energy loss associated with full revolutions of the particle, $2\pi Rf = 1250\unit{\kT}$. To investigate the FT, rare events with negative entropy production have to be sampled with high accuracy. This constrains the maximal trajectory length $t$ to approximately $2\unit{s}$ and the range within the FT can be tested to $\pm3$. Figure~\ref{fig2}(b) shows this section of the black (closed bars) histogram around $\sred=0$. The excellent agreement between the logarithm of the probability ratio $p(\sred)/p(-\sred)$, black squares in Fig.~\ref{fig2}(c), and the black dashed line with a slope of $1$ confirms the validity of Eq.~\eqref{eq:FT} for uncoupled states. The red histogram (open bars) in Figs.~\ref{fig2}(a) and \ref{fig2}(b) demonstrates the situation for coupled states. Most prominent is the enhanced probability at $\sred = 0$. Since the red dots in Fig.~\ref{fig2}(c) do not agree with the dashed line of slope $1$ this apparent entropy production does not obey the FT. Rather a linear relation as given by Eq.\eqref{eq:FT} with $\alpha \simeq 0.65$ is found. \begin{figure} \includegraphics[width=\linewidth]{fig3} \caption{(color online) (a) Slope $\alpha$ vs plasma parameter $\Gamma$ for $t = 1.75\unit{s}$. (b) Slope $\alpha$ for different trajectory lengths $t$. The black squares correspond to $\Gamma = 0$ and the red dots to $\Gamma = 300$. The deviation of the black squares from $\alpha = 1$ (black dashed line) determines the statistical errors to be less than $5\%$. \label{fig3}} \end{figure} In additional experiments, we observe linear relations according to Eq.~\eqref{eq:FT} with different slopes $\alpha$, which depend on two parameters: (i) the plasma parameter $\Gamma$, and (ii) the trajectory length $t$, as shown in Figs.~\ref{fig3}(a) and \ref{fig3}(b). Clearly, the FT is confirmed for arbitrary trajectory lengths in uncoupled states (black squares in Fig.~\ref{fig3}(b)). The obvious dependence of $\alpha$ on $\Gamma$ resembles the transition from an uncoupled to a coupled state. For $\Gamma = 300$, $\alpha$ decays with increasing length $t$, from $1$ to $0.65$. A similar time dependence has been also observed in~\cite{haya10}; however, there it was not identified as an inherent feature of hidden degrees of freedom. In additional measurements performed for coupling two different NESS, we also found such a linear relation. Therefore, we exclude symmetry as the sole origin of this behavior. {\it Discussion.}---First, we explain why for $t \rightarrow 0$ the slope $\alpha$ approaches $1$. In general, deviations from the FT must be caused by the interaction with the hidden particle. In a short time expansion to lowest order in $t$, we can neglect changes in the interaction force during the motion of the observed particle. Thus, the interaction force entering Eq.~\eqref{eq:stot_red} through $\nucg$ becomes constant in this limit. The apparent entropy production then becomes equivalent to that of an effective one-particle system subject to a Markovian dynamics with mean local velocity $\nucg$, which trivially obeys the FT. This effective description is valid only to lowest order in time since taking into account higher order terms would include contributions arising from the correlated motion of the observed and the hidden particle. % ---------- Theory ---------- Although we are able to qualitatively understand the influence of coupling it remains a surprising feature why, in all of our experiments presented so far, only the slope of the FT is affected by the coupling strength while the linear relation Eq.~\eqref{eq:FT} remains untouched. In order to elucidate this result we define the function \begin{equation} \label{eq:function} f(\sred) \equiv \ln\left[p(\sred)/p(-\sred)\right], \end{equation} which we assume to be analytic. First, we note that $f$ is antisymmetric by construction, and thus for small entropy productions, $\sred \ll 1$, $f$ trivially must be linear up to corrections of third order or higher~\cite{pugl05}. Second, we discuss $f$ for large entropy productions, $\sred \gg 1$. Solving Eq.~\eqref{eq:function} for $p(-\sred)$ and integrating over all $\sred$ yields \begin{equation} \label{eq:norm} \int_{-\infty}^{+\infty}p(\sred)\,e^{-f(\sred)}\,d\sred = 1, \end{equation} by normalization. We assume that $p(\sred)$ does not decay faster than a Gaussian as we have observed in all our measurements. For any quantity consisting of independent contributions the central limit theorem would imply a Gaussian. Any correlation will typically lead to an even slower decay. Under this assumption convergence of the integral in Eq.~\eqref{eq:norm} requires that $f(\sred) = \LandauO{\sred^2}$. Since, in addition, $f$ is antisymmetric we expect the generic asymptotic behavior to be linear, $f(\sred) \sim \sred$, with a slope generally different from the one for small $\sred$. Summarizing these arguments, we expect a linear function both for small and for large entropy production for any time $t$. For intermediate entropy production this reasoning leaves the possibility of a nonlinear behavior. Even though we have found a constant slope for most experimental parameters, by fine-tuning the system and plasma parameter, we can observe an obviously nonlinear result, as shown in Fig.~\ref{fig4}(a). % ---------- Experiment II ---------- \begin{figure} \includegraphics[width=\linewidth]{fig4} \caption{(color online) (a)~$\ln\left[p(\sred)/p(-\sred)\right]$ as a function of $\sred$ for $t = 3\unit{s}$. (b)~Typical trajectories of the observed (blue/thick line) and hidden particle (gray/thin line). \label{fig4}}

48 — 1205.0677

\caption{(De)confinement (D/C) properties for cosmological \blue{Liu-Tseytlin} \cite{EGM} and \red{AdS/Schwarzschild} (i.e. constant dilaton) backgrounds \cite{Tetradis}. The stated properties hold for all values of $a_0$ attained during cosmological evolution unless stated otherwise. The value of $a_{0,\ast}$ is given in \eqref{a0star}.}

49 — 1205.1286

\caption{\label{figure04} Spatially-resolved scan of the time-resolved photoluminescence signal. The different panels show the fitted amplitude $A_{slow}$ and decay time $\tau_{slow}$ of the slow decay component and the amplitude $A_{fast}$ and decay time $\tau_{fast}$ of the fast decay component. \protect\includegraphics[height=3.7mm]{A}, \protect\includegraphics[height=3.7mm]{B} and \protect\includegraphics[height=3.7mm]{C} present examples of decay transients of the spatially-resolved scan recorded at the position \protect\includegraphics[height=3.7mm]{A} the bulk material, \protect\includegraphics[height=3.7mm]{B} the photonic crystal membrane, and \protect\includegraphics[height=3.7mm]{C} the photonic crystal waveguide.}

\caption{Overview on the extracted decay times and amplitudes for \protect\includegraphics[height=3.7mm]{A} the bulk material, \protect\includegraphics[height=3.7mm]{B} the photonic crystal membrane, and \protect\includegraphics[height=3.7mm]{C} the photonic crystal waveguide.}

50 — 1205.2344

\caption{ \textbf{Left, middle:} Light curves and SEDs for the quiescent state of the EC/BLR model, using a blackbody approximation for the BLR spectrum. The histograms in both figures are the results of our simulation, with the energy band chosen shown in the legend of the left figure. The SED snapshots times are shown in legend of figures and marked with matching color segments in the light curves plot. In this and all the paper SED plots the data points are multiwavelength SEDs of PKS~1510$-$089 mostly in the spring of 2009 from \citet{abdo_etal:2010:pks1510_multiwavelength_flares_2008_2009} and \citet{dammando_etal:2011:pks1510_with_agile_march_2009}. The optical/infrared points, also show in the inset, are for the following dates: blue for March 10, orange for March 18, black for March 19, red for Match 25/26 (flare peak). % The \gray, \fermi/LAT spectra are: blue squares for the quiescent state of the early 2009 \citep[see][]{abdo_etal:2010:pks1510_multiwavelength_flares_2008_2009}, red squares the end of March 2009 flare, black empty squares an earlier weaker flare (flare `a' in \citealp{abdo_etal:2010:pks1510_multiwavelength_flares_2008_2009}) which we show as a plausible reference for the gamma-ray state before and after the March 2009 flare. % The gray points in radio, submm and \xray are not strictly simultaneous but we regard them as representative because of the very modest variability exhibited by the source during those months in these bands. The \xray data include the XRT data averaged during the March 2009 flare, and five year integrated BAT data in hard \xrays. %% \textbf{Right:} The SEDs for the quiescent state of the BLR model using the \citet{tavecchio_gg:2008:blr_and_blazars} BLR spectrum (parameters are listed in Table~\ref{tab:model_parameters}, \texttt{nf/blr}). \label{fig:blr15nf_tave} \label{fig:blr15nf_bb} }

51 — 1205.2813

\caption{(Color online) 3D contour plot of density $|\phi|^2$ of a ring-shaped dipolar $^{164}$Dy BEC for $ a_{dd}=130a_0, N=5000, a=120a_0 $ and (a) $\alpha=0$, (b)$\alpha=\pi/6$, (c) $\alpha=\pi/3$, (d) $\alpha=\pi/2$. The density at the contour is 0.005. {\color{red} The variables $x$, $y$, and $z$ are in units of $l_0 (= 1$ $\mu$m.)} }

\caption{(Color online) 3D contour plot of density $|\phi|^2$ for a shell-shaped dipolar BEC with $ g_{dd}=100, $ and (a) $g=300$, (b)$g=600$, (c) $g=1000$, (d) $g=2000.$ The density on the contour is 0.0005. {\color{red} The variables $x$, $y$, and $z$ are in units of $l_0 (= 1$ $\mu$m.)} }

52 — 1205.3176

\caption{Mass specific frequency of metal-poor GCs for cluster and field dIrrs, and Virgo dEs -- including both nucleated and non-nucleated objects, as well as systems with disc features and central star formation. Galaxies having no blue GCs have been assigned a \tnblue\$=0.12$, while diagonal lines correspond to the loci of systems with 1, 10 and 100 blue GCs. The shaded region indicates, as in Fig.\,\ref{fig:harass_lsb}, the potential location of harassed disc-like galaxies. Late-type \emph{field} dwarfs preferentially have lower specific frequencies than both late- and early-type \emph{cluster} dwarfs -- a feature that is more prominent at the high-mass end and holds even for nucleated dIrrs. The arrow indicates the effect of \mstar\doubling had the dE progenitors not had their star formation quenched (see text for details).}

\caption{Statistical properties of the metal-poor GCSs in the different \mstar\$>2\times10^{8}$ \msun\dwarf types in Fig.\,\ref{fig:dwarfs_gcs}. Each non-diagonal element in the upper table indicates the confidence level, according to a \ks\test, with which we can rule out the null hypothesis that the two\tnblue\distributions are drawn from the same parent population. The diagonal elements show the fraction of galaxies of each type that have ten or more blue GCs. In the lower table we show the corresponding lower quartiles, medians and upper quartiles of the\tnblue\distributions.}

\caption{Fraction of red (metal-rich) GCs as a function of host galaxy stellar mass. Galaxies consistent with having no red clusters have been slightly offset for displaying purposes. The gray curve is a running average, and shows that there is a trend for \mstar\$\lesssim 10^{11}$ \msun\galaxies to have lower metal-rich GC proportions. Blue curves indicate the expected evolution of a$10^{10}$ \msun\galaxy if mass stripping occurs from the outside-in (see text for details). In both cases the galaxy hosts a GCS with subpopulations having an identical spatial distribution to that of the MW, and initial red fractions of$\approx$\,30\% and $\approx$\,15\%. Contrary to observations, the red GC fraction would tend to increase towards lower masses -- their less extended spatial distribution makes them more resilient to stripping than the blue subpopulation.}

\caption{Empty circles show \tnblue\for ACSVCS Virgo dwarfs as a function of their tridimensional distance to M87, while the gray curve represents a smoothed version of the radial trend. Filled circles indicate the final fraction of stellar mass for the harassed galaxies simulated by\citet{Mastropietro2005} as a function of their average clustercentric radius (see text for details). Galaxies that spend longer times deep within the cluster potential are affected the most. The trend of innermost dEs preferentially having richer GCSs than the outermost ones is the opposite than predicted by mass-stripping mechanisms. }

53 — 1205.3303

\caption{ \label{Fig_mag} (Colour online) Time evolution of the local magnetization $m_l\left(t\right)$ after the six quench protocols in Fig.\ref{Fig_phase_diagram}. Free fermion (SC theory) results are indicated by full (broken) lines (${\color{black}-}\,l=32$, ${\color{red}-}\,l=64$, ${\color{green}-}\,l=96$, ${\color{blue}-}\,l=128$). } \end{figure} We have calculated the time evolution of the local magnetization $m_l\left(t\right)$ for finite open chains of length $L=256$ at the positions $l=32,\,64,\,96$ and $128$. In Fig. \ref{Fig_mag} the exact free fermion results are compared with the predictions of the modified SC theory for the six different quench protocols indicated in Fig. \ref{Fig_phase_diagram}. If the quench is performed between two paramter points within the ordered phase, the modified SC theory represents an excellent description of the relaxation process, in particular for short times ($t<T_{\rm period}/2$). If the quench involves the disordered phase as well or the XX point the agreement is less good. However, the qualitative features of the time evolution of the magnetization is similar in all cases in Fig. \ref{Fig_mag} and it has the same form as found previously for the transverse Ising chain\cite{Igloi_11,Rieger_11}. The exponentially decreasing initial part of $m_l\left(t\right)$ is followed by a quasi-stationary plateau and afterwards there is an exponentially increasing reconstruction part. Then the process is repeated quasiperiodically. In the $L \to \infty$ limit the local magnetization is given by % \beqn m_l\left(t\right)=m_l\left(0\right) \exp\left(-t\frac{2}{\pi}\int_0^{\pi} dp \,v_p \tilde{f}_p\theta\left(l-v_p t\right)\right.\nonumber\\ \left.-l\frac{2}{\pi}\int_0^{\pi} dp \, \tilde{f}_p \theta\left(v_p t-l\right) \right) \label{m_l} \eeqn % in the modified SC theory, which defines the quench-dependent relaxation time $\tau_{\rm mag}$ and the correlation length $\xi_{\rm mag}$ as % \be \tau_{\rm mag}^{-1}=\frac{2}{\pi}\int_0^{\pi} dp \,v_p \tilde{f}_p, \quad \xi_{\rm mag}^{-1}=\frac{2}{\pi}\int_0^{\pi} dp \, \tilde{f}_p. \label{tau_mag} \ee % \subsection{Correlation function} \begin{figure} \begin{center} \includegraphics[width=0.45\textwidth]{fig4.eps} \end{center} \caption { \label{Fig_eqtc} (Colour online) Equal-time correlation function $C_t^{xx}\left(r\right)$ for fixed time $t$ after the six quench protocols in Fig.\ref{Fig_phase_diagram} as a function of the distance $r$. Free fermion (SC theory) results are indicated by full (broken) lines (${\color{black}-}\,t=10$, ${\color{red}-}\,t=20$, ${\color{green}-}\,t=40$, ${\color{blue}-}\,t=60$, ${\color{magenta}-}\,t=80$, ${\color{cyan}-}\,t=100$). }

54 — 1205.4165

\caption{\label{fig:endToEndGrowthMucaLowDens}(Color online) (a) End-to-end distribution function for site occupation $p=0$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackSolid.pdf}, black solid), $0.13$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.25$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.38$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $0.51$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black) for increasing peak height. The curves are interpolating lines through the data (whose markers have been omitted for better visibility). The results of the two algorithms agree within the line thickness. The influence of the disk diameter $\sigma_i$ is negligible in this density regime and chosen here to be $\sigma_2$. (b) and (c) are the corresponding plots for the tangent-tangent correlations and the mean square end-to-end distance (in units of squared bond length $b^2$).}

\caption{\label{fig:endToEndGrowthMucaHighDens}(Color online) End-to-end distribution function for site occupation $p=0.64$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.76$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $1$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black). The black solid curve is the end-to-end distribution of the free polymer as reference. The data marked by $\circ$ are from the growth algorithm and $+$ are from the multicanonical algorithm. The different plots are made for $\sigma_{1,2,3}$. The inset shows in each case the regime $p=0.64, 0.76, 0.89$ (the black solid curve is again the case $p=0$ as reference), where both the influence of the low-density regime and the influence of the small cavities play a role.}

\caption{\label{fig:tanTanCorrAllDens}(Color online) Tangent-tangent correlations for $p=0.64$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.76$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $1$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black). The black solid curve is for $p=0$ as reference. The data marked by $\circ$ are from the growth algorithm, whereas $+$ come from the multicanonical algorithm. (a), (b), and (c) differ in the disk diameter. The larger the disk diameter the stronger the anti-correlations on short length scales.}

\caption{\label{fig:meanSquaredEndToEnd}(Color online) Mean square end-to-end distance (in units of squared bond length $b^2$) for $p=0.64$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.76$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $1$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black) from top to bottom. The data marked by $\circ$ are from the growth algorithm; $+$ come from the multicanonical algorithm. The black solid curve shows the free polymer case which scales as $[\langle R_{\mathrm{ee}}^2 \rangle] \propto N$.}

55 — 1205.4189

\caption{New {\it conventional} states in the $c\bar{c}$, $b\bar{c}$, and $b\bar{b}$ regions, ordered by mass. Masses $m$ and widths $\Gamma$ represent the weighted averages from the listed sources. Quoted uncertainties reflect quadrature summation from individual experiments. In the Process column, the decay mode of the new state claimed is indicated in parentheses. Ellipses (...) indicate inclusively selected event topologies; \ie additional particles not required by the Experiments to be present. A question mark (?) indicates an unmeasured value. For each Experiment a citation is given, as well as the statistical significance in number of standard deviations (\#$\sigma$), or ``(np)'' for ``not provided''. The Year column gives the date of first measurement cited. The Status column indicates that the state has been observed by at most one ({\color{red} NC!}-needs confirmation) or at least two independent experiments with significance of $>$5$\sigma$ (OK). The state labelled $\chi_{c2}(2P)$ has previously been called $Z(3930)$. See also the reviews in \cite{Brambilla:2004wf,Eichten:2007qx,Eidelman:2008zzc,Godfrey:2008nc,Barnes:2009zza,Pakhlova:2010zz,Brambilla:2010cs}. Adapted from \cite{Brambilla:2010cs} with kind permission, copyright (2011), Springer. }

56 — 1205.4286

\caption{Panel (a): experimentally determined excitation fraction in Ba$^{+}$ versus theoretical predictions of a density matrix ($\bullet$) and rate equation (\textcolor{red}{$\circ$}) treatment. Panel (b): experimentally determined excitation fraction in Ca versus predictions of a 4-level rate equation model. In both panels the solid line has a unit slope and represents perfect agreement between experiment and theory.}

\caption{Measured rate constant as a function of Ba$^+$ population fractions. Experimental result in black. Best fit to the four entrance channel model shown as open red circles (\textcolor{red}{$\circ$}). %The red error bars are representative of the best fit's sensitivity to large variation ($0 - 10^{-9}$cm$^{3}$Hz) in the channels with poor sensitivity. }

\caption{Measured rate constant as a function of Ca population fractions. Experimental result in black. Best fit to the four entrance channel model shown as open red circles(\textcolor{red}{$\circ$}). %The red error bars are representative of the best fit's sensitivity to large variation ($0 - 10^{-9}$cm$^{3}$Hz) in the channels with poor sensitivity. }

57 — 1205.5510

\caption{ SEDs accumulated during the 2010 November flare (in colors, data from~\cite{Vercellone2011:ApJL_3C454_nov2010}) compared with a SED accumulated during a particularly low \gray{} state in Fall 2008 (in black, data from~\cite{Vercellone2010ApJ_P3}). }

58 — 1205.6941

\caption{(color online). Leap-frog staggered-grid scheme: The left part shows the time-stepping where \textcolor{blue}{1)} the new $u$ components \textcolor{blue}{(blue)} are computed by the previous $u$ and the spatial differences of old $v$-values. \textcolor{red}{2)} Then (knowing $u$ at $t_{n+1}$) the new $v$ at $t_{n+1}$ \textcolor{red}{(red)} are computed. The mass $m$ and potential $V$ enter the scheme in a Crank-Nicolson-type time averaging over the current- and previous-time values. The right part of this figure shows the scheme for the spatial derivatives.}

59 — 1206.0193

\caption{\coloron The Cooper-pair sluice. (a) Equivalent circuit of the sluice and scheme of the measurement set-up. (b) False-color scanning-electron micrograph of a representative device. (c) Time evolution of the control parameters for a typical pumping cycle. % The gate is used as a piston to change the number of Cooper-pairs on the island, % the SQUIDs as valves so as to impart direction to the flow of charge. }

\caption{\coloron %(a) Representative radiofrequency pulses applied to the control parameters in order to achieve pumping. Single Cooper-pair pumping. Pumped charge $Q_p$ versus peak-to-peak amplitude $\dng$ and offset $n_g^0$ of the gate drive. Dashed lines enclose the diamond-shaped regions where $Q_p/2e$ is expected to be constant and quantized (assuming ideal operation).}

\caption{\coloron Crossover between pure Cooper-pair and mixed Cooper-pair-quasiparticle dynamics. (a) Pumped charge $Q_p$ versus gate offset $n_g^0$ for increasing gate amplitudes $\dng$ (bottom to top). (b) Circles: same data as in (a), projected on the $n_g^0$ axis and plotted versus $\dng$. Dashed line: asymptotic adiabatic-limit expectation. In both panels, the first three pumping plateaus are indicated by arrows. }

\caption{\coloron Nonadiabatic pumping and Landau-Zener transitions. (a) Instantaneous energies of the ground and first excited state versus time for a pumping period. In the model described in the text, nonadiabatic transitions are localized at level crossings, and occur with probability $P_{LZ}$. Decoherence induces complete dephasing between subsequent transitions, and relaxation via inelastic Cooper-pair tunnelling (both are indicated by wavy arrows). (b) Normalized pumped charge $Q_p/Q_0$ versus Josephson coupling $J_{max}$, for a set of frequencies in the range of 70 and 120 MHz. The experimental traces (squares) are vertically offset by 0.2, and plotted together with the best-fit of Eq.~\ref{eq:easyqp} to them (solid lines). (c,d) Asymptotic pumped charge $Q_0$ versus $f$ (c) and $\lambda$ parameter versus $1/f$ (d) for three different measurement sets (squares, circles, and triangles). The solid line in (d) is a fit to the data of the expression $\lambda=f_0/f$.}

60 — 1206.0194

\caption{\label{fig1} Steady-state spatial distribution for the dipolar exciton BEC density in a trap for the low-frequency driving at $R_0=0.1$. %The colorbar shows the exciton density $|\Psi(\bm{r})|^2$. Inset shows the evolution of the exciton density profile with rising the pumping rate $R_0$: 0.1 ({\color{red} $\blacksquare$}), 0.3 ({\color{darkgreen} \Large $\bullet$}), and 0.5 ({\color{blue} $\blacktriangle$}). The curves represent the Thomas-Fermi distribution function.}

61 — 1206.0412

\caption{ {\bf Orbital selection scenario: } Dependence of a) the iron $3d$ subspace occupancy $n_d$ and b) the effective spin quantum number $s$ on the Hund's coupling $J$, for both unligated and ligated heme models. The physically relevant region \cred $0.5~$eV~$<J<1$~eV \cblack is highlighted in yellow. Isosurfaces of the real-space representation of the electronic spectral density of the HOMO of FeP-d for c) $J=0$~eV and d) $J=0.8$~eV. The large central sphere shows the location of the iron atom, and the four blue spheres indicate nitrogen atoms.}

62 — 1206.1086

\caption{Properties of models at core bounce, where bounce is defined as the maximum compression of the central density during the launching of the bounce shock. Models shown are: \base\(black; all opacities), \basenes\(blue; without NES) with the other NIS opacity variation models discussed in Section~\ref{sec:basenis} indistinguishable from models \base\and\basenes\at bounce and omitted for clarity; the NIS opacity variation models with the IPA EC from Section~\ref{sec:ipanis}, \ipa\(orange; all NIS opacities), \ipanis\(red; without NIS opacities; no NES, no NPS, nucleon IS), \ipanes\(green; without NES), but not \ipareds\(nucleon IS), which is indistinguishable from model \ipa\at bounce and omitted for clarity. The pair opacity test models (Section~\ref{sec:pair}) and improved nucleon EC model (Section~\ref{sec:abem}) are also indistinguishable from model \base\and omitted for clarity. The panels are: radial velocity (upper left), density (upper center), entropy (upper right), temperature ($kT$, lower left), net electron (or proton) fraction ($Y_e$, lower center, solid lines), net lepton fraction ($Y_L = Y_e + (n_{\nu_e} - n_{\bar{\nu}_e})/n_{\rm baryons}$, lower center, dashed lines), and pressure (lower right). All quantities are plotted relative to enclosed rest-mass in \msun. \label{fig:bounce}}

\caption{Shock trajectories in km, versus time after bounce, for all models in Section \ref{sec:basenis}. The models plotted are \base\(black; all opacities included), \basenis\(red; without NIS opacities; no NES, no NPS, nucleon IS), \basenes\(green; without NES), and \basereds\(blue; nucleon IS). Shock position is computed by bisecting the pair of mass shells with the largest negative radial velocity gradient $-\partial v_r/\partial r$. \label{fig:baseshock} }

\caption{Shock trajectories in km, versus time after bounce, for all models in Section \ref{sec:ipanis}. Models are \ipa\(orange; all NIS opacities), \ipanis\(red; without NIS opacities; no NES, no NPS, nucleon IS), \ipanes\(green; without NES), and \ipareds\(blue; nucleon IS). \label{fig:ipashock}}

\caption{Net heating rates in $\mbox{ergs g}^{-1}\,{\rm s}^{-1}$ for models \base\(black) and \basepair\(red) at 30~ms after bounce. Line styles indicate net heating by: emission/absorption by \nue\(solid) and \nuebar\(dotted), by \numt\numtbar\pair sources (dashed) and by NIS of\numt\numtbar\(dash-dotted). Net pair heating for \nue\nuebar\is approximately one order of magnitude or more smaller than the\numt\numtbar\counterpart and is omitted from this plot for clarity.\label{fig:pairheat} }

63 — 1206.1304

\caption{[COLOR ONLINE] \textit{Ab initio} X~$^2\Sigma_{u}^+$ (in black), B~$^2\Sigma_{g}^+$ (in red) and A~$^2\Pi_{u}$ (in blue) states of Ca$_{2}^+$. Dashed lines show calculation with a pseudopotential, while solid lines show the results of an all electron correlated calculation. For the A~$^2\Pi_{u}$ state, results of all electron MRCI calculations are shown in green \color{green4}$\bullet$\color{black}. Note that a$_0$ is the Bohr radius (atomic unit of length).}

\caption{[COLOR ONLINE] (a) Calculated radiative lifetimes of bound levels of $^{40}$Ca$_2^+$ in the B~$^2\Sigma_{g}^+$ state, on a log-scale. The shorter lifetimes (blue \color{blue}$\bullet$\color{black}) correspond to bound levels localized in the inner well, the longer lifetimes (red \color{red}$\circ$\color{black}) to levels localized in outer well, and the increasingly longer lifetimes (green \color{green4}$\Diamondblack$\color{black}) to levels spread over both wells. (b) show examples of each cases: $v$=0 and 7 in the inner well, $v$=6 in the outer well, and $v$=40 in both. (c) depicts $|\psi|^2$ of $v$=6 on a log-scale; showing that the amplitude in the inner well is extremely small. The reverse is true for $v$=7 and so on.}

64 — 1206.1617

\caption{\small Comparison of the particle's signal $ \vec a_\text{SP}(t)$ (\Cred{$\bullet$}) to the theoretical curves based on the position sensor (---). $a_y$ points with the arm and $a_x$ measures the force in direction of the movement. No force is exerted along the $z-$axis (the green lines represent the uncertainty of the calibration). Note, that acceleration is measured in $g=9.8\,\metre\per\second\squared$ . }

\caption{\small A sample trajectory of the instrumented particle seen by the camera (\Cblue{---}) or smartPART (\Cgreen{$\bullet$}), it is $\fprop=3\,\hertz$. The absolute orientation enables us to re-express the camera measurement of the particle (lab frame) in the moving frame of the particle and vice-versa. In the former gravity is subtracted and in the particle frame gravity is represented by the \Cred{red} line. }

65 — 1206.2748

\caption{ \newline \textbf{Left:} Linear growth-rates of a tearing-mode in a circular plasma with concentric ideal wall are compared between \inred{CASTOR using its own vacuum field module, CASTOR coupled with STARWALL, and JOREK coupled with STARWALL}. Very good agreement is observed for a variety of different plasma resistivities and plasma-wall distances. \newline \textbf{Right:} The current potential distribution (arbitrary units) reflecting the $2/1$ tearing mode structure is plotted during the linear phase ($\eta_\text{Jorek}=1\cdot10^{-6}$, $r_\text{w}=1.2$).}

66 — 1206.2757

\caption{(Color online) Three nodes schematic diagram illustrating impact of delay. Arrows depict the direction of information flow as governed by Eq.(\ref{cml}). The dashed lines show the flow of information from the $(t-2)^{th}$ time step. For $\tau=0$, evolution of all nodes ({\color{orange}{$\bullet$}}) receive information from the second node (left panel), whereas in presence of delay, evolution of connected nodes at a particular time do not involve any common term (right panel). For both panels, first and third nodes are connected with the second one leading to the construction of the smallest possible bipartite network.}

67 — 1206.3353

\caption{\label{F4} Comparison of the two probabilities.~ \dotted ~line corresponds to the standard one $p_{l}=\rme^{-\beta E_{l}}$, and \broken ~line to $p_{l}=f(\beta E_{l})$}

\caption{\label{F1} Entropies as function of $\Omega$. \dotted ~ and ~ \broken ~lines correspond to $S_{\beta}/k$ and $S/k$ respectively ($p_{l}=1/\Omega$ equipartition)}

\caption{\label{F2} Entropies as function of $\Omega$. \dotted ~and~ \broken ~ lines correspond to $S_{\beta}/k$ and $S/k$ respectively ($p_{l}=1/\Omega$ equipartition)}

68 — 1206.3839

\caption{Distribution of the arrival directions of 69 UHECR, represented by blue circles ({\color{blue}$\circ$}), with energy $E\geq5.5\times10^{19}\,{\rm eV}$ reported by PAO in 2010, in the equatorial coordinates plotted using the Hammer projection. The solid red line represents the exposure boundary of the PAO experiment. The locations of Centaurus A (${\color{black}\blacksquare}$) and M87 (${\color{black}\blacklozenge}$) are also shown for reference.}

69 — 1206.6707

\caption{\label{fig:fig3-model} (a) Cartoon of graphene on quartz substrate (GOQ). (b) Schematic band diagram from Density functional theory (DFT) of graphene on quartz (GOQ) assuming no doping as in free-standing case. Possible optical transitions are shown where all states above the Dirac point are accessible. (c) Schematic energy values that represent the peak position in the optical transitions for GOQ. All dashed lines and arrows are representing theoretical predictions. The brown dashed line(\color{brown}{---}\color{black}) at 4.11 eV represents the result from local density approximation (LDA) theory\cite{YangNanL} which accounts only for band to band transitions. The pink dashed line (\color{pink}{---}\color{black}) at 5.20 eV represents the result from GW calculations which takes into account many-body electron –electron (e-e) interactions\cite{YangNanL}. The blue arrow (\color{blue}{$\rightarrow$}\color{black}) indicates the difference in energy of GW from LDA calculations. The green dashed line (\color{green}{---}\color{black}) at 4.53 eV represents the energy value predicted by GW- Bethe Salpeter Equation (GW-BSE) approach which includes also the electron –hole (e-h) interactions\cite{YangNanL}. The red arrow (\color{red}{$\rightarrow$}\color{black} ) indicates the energy difference between GW and GWBSE calculations which is the contribution only from e-h interactions. The thick purple line (\color{purple}{---}\color{black}) at 4.60 eV is representing the asymmetric peak position in our experimental result described in text for GOQ. The thickness is proportional to the error bar which is affected mostly by the fitting procedure.The closeness of our result with the GW-BSE prediction indicates the presence of both e-e and e-h interactions in GOQ.}

\caption{\label{fig:fig4-model} (a) Cartoon of graphene on copper substrate (GOC). (b) Schematic band diagram from Density functional theory (DFT) of graphene on copper (GOC) assuming considerable electron doping based on DFT calculations\cite{GiovannettiPRL}. Possible transitions are shown with cyan arrows ( \color{cyan}{$\rightarrow$}\color{black} ) to states above the Fermi level. States below the Fermi level but above the Dirac point are not accessible (represented by crossed cyan arrow). (c) Schematic energy values that represent the peak positions in the optical transitions for GOC. All dashed lines and arrows are representing theoretical predictions where electron doped graphene (0.01 electrons per unit cell of graphene) is considered. The brown dashed line (\color{brown}{---}\color{black}) at 4.11 eV represents the result from local density approximation (LDA) theory which accounts only for band to band transitions\cite{YangNanL}. The pink dashed line (\color{pink}{---}\color{black}) at 4.91eV represents the result from GW calculations which takes into account many-body e-e interactions\cite{YangNanL}. The blue arrow ( \color{blue}{$\rightarrow$}\color{black} ) indicates the difference in energy of GW from LDA calculations. The green dashed line (\color{green}{---}\color{black}) at 4.49 eV represents the energy value predicted by GW- Bethe Salpeter Equation (GW-BSE) approach which includes also the e-h interactions\cite{YangNanL}. The red arrow (\color{red}{$\rightarrow$}\color{black} ) indicates the energy difference between GW and GWBSE calculations which is the contribution only from e-h interactions. The thick purple line (\color{purple}{---}\color{black}) at 4.96 eV is representing the symmetric peak position in $\sigma_{1}(\omega)$ of our experimental result described in text for GOC. The thickness is proportional to the error bar which is affected mostly by fitting procedure. The fact that our experimental peak position is close to GW prediction indicates that e-h interactions are screened almost fully}

70 — 1206.6802

\caption{\label{fig:continous_temp}Peak heights versus normalized packing fraction. $g_{m}-1$ is plotted versus $\phi/\phi^{*}$ for the first four peaks, $m= 1,...,4$. $\square$ are extracted from data where the volume fraction is continuously ramped to higher values. {\color{red}$\bullet$} is extracted from data quenched to $\phi$. At all volume fractions the vertical error bars are on the order of the marker size. For $\phi > \phi^{*}$ the data points are tightly spaced. }

71 — 1207.0110

\caption{{\bf Critical strain} (color online): \leg{(a):} $\gamma_c$ at the onset of fingering as a function of $b/D_0$ for Setup1 (\textcolor{blue}{$\blacksquare $}) and Setup 2 (\textcolor{red}{o}). \leg{(b):} Setup 2- critical strain $\gamma _c$,$(\textcolor{red}{\circ})$, and strain to fracture $\gamma _R, (\textcolor{blue}{\square})$, as a function of the shear modulus $G$, $D_0=23mm$.}

72 — 1207.0519

\caption{(Color online) Thermal evolution of the $ac$ susceptibility for the $x=0.4$ sample, taken at $h_{ac}=$10 Oe, and various frequencies. (a)\$\chi_{ac}^{\prime\prime}$: the out-of-phase component; (b) $\chi _{ac}^{\prime}$: the in-phase component. The vertical dashed line emphasizes the absence of\a frequency-dependent\variation in the peak position (though an intensity reduction is evident). Inset: an expansion of the lower peaks showing a 10\% reduction in the peak position for four decade frequency variation (denoted by the short vertical arrows).}

73 — 1207.1571

\caption{ (Color online) Validation cases: a) lid-driven cavity ($10^6$ cells), b) transient Poiseuille flow, c) coronary artery. The dotted (online: red) line is the cross-section along which the results are evaluated. Results: d,f) velocity components $u$, $v$, $w$ (parallel to the $x$, $y$ and $z$ axis, respectively) along the cross-section; e) velocity $u$ at several times (in seconds); g,h,i) pressure along the cross-section. The results obtained on the CPU and GPU are shown as symbols (\textcolor{red}{$\bullet$}) and solid lines, respectively. Only part of the available CPU data are shown for clarity.%better visualization. \label{fig:res}}

74 — 1207.1737

\caption{(color online) A visual summary of our results. Top left: Beyond critical density and activity levels the active colloidal fluid separates into dense and dilute phases. The clusters coarsen over time (see \ref*{Supplement-fig:coarsening-movie} in \cite{SupplementalInformation}). Top right: The static structure factor $S(\boldsymbol{k}) = \frac{1}{N} \left\langle \sum_{ij} e^{i \boldsymbol{k} \cdot \boldsymbol{r}_{ij}} \right\rangle$, restricted to the interiors of large clusters. These signatures resemble those of a high temperature colloidal crystal near the crystal-hexatic phase transition. Bottom left: A heat map of the pressure in the active solid material. It is heterogeneous and highly dynamic, indicating that external stresses would produce a complex response. Bottom right: Log-log plot of the mean square displacement of a tagged particle in the active solid. At intermediate time scales, it exhibits anomalous superdiffusive transport.} \label{fig:summary} \end{figure} In this letter we explore a minimal active fluid model: a system of self-propelled smooth spheres interacting by excluded volume alone and confined to two dimensions. Unlike self-propelled rods \cite{PhysRevE.77.011920, PhysRevLett.101.268101, PhysRevLett.106.128101, PhysRevE.82.031904, C2SM06960A}, these particles cannot interchange angular momentum and thus lack a mutual alignment mechanism. Recent simulation and experimental studies have shown that this system exhibits giant number fluctuations \cite{PhysRevLett.108.235702} and athermal phase separation \cite{PhysRevLett.108.235702,2012arXiv1202.6264T} that are characteristic of active fluids \cite{0295-5075-62-2-196, PhysRevLett.97.090602, Cates2010}. Here we employ extensive Brownian dynamics simulations to characterize the phase diagram of this system and we develop an analytic model that captures its essential features. We show that this nonequilibrium system undergoes a continuous phase transition, analogous to that of equilibrium systems with attractive interactions, and that the phase separation kinetics demonstrate equilibrium-like coarsening. These structural and dynamic signatures of phase separation and coexistence enable an unequivocal definition of phases in this nonequilibrium, active system. Finally, we find that the dense phase is a dynamic new form of material that we call an ``active solid''. This material exhibits structural properties consistent with a 2D colloidal crystal near the crystal-hexatic transition point \cite{PhysRevLett.104.205703, PhysRevE.77.041406}, but is characterized by such anomalous features as superdiffusive transport at intermediate timescales and a heterogeneous and dynamic stress distribution (see Fig.~(\ref{fig:summary})). \noindent \emph{Model and Simulation Method}: Our system consists of smooth spheres immersed in a solvent and confined to a plane, similar to experimental systems of self-propelled colloids sedimented at an interface \cite{2012arXiv1202.6264T}. Each particle is self-propelled with a constant force, and interactions between particles result from isotropic excluded-volume repulsion only. We include no mechanism for explicit alignment or transmission of torques between particles. The state of the system is represented by the positions and self-propulsion directions $\{\boldsymbol{r}_i, \theta_i\}_{i=1}^N$ of all particles. Their evolution is governed by the coupled overdamped Langevin equations: \begin{align} \dot{\boldsymbol{r}}_i &= D \beta \left[ \boldsymbol{F}_{\mathrm{ex}}(\{\boldsymbol{r}_i\}) + \Fp \hat{\boldsymbol{\nu}}_i \right] + \sqrt{2 D} \, \boldsymbol{\eta}^T_i \\ \dot{\theta}_i &= \sqrt{2 \Dr} \, \eta^R_i \end{align} Here $\boldsymbol{F}_{\mathrm{ex}}$ is an excluded-volume repulsive force given by the WCA potential $V_{\mathrm{ex}} = 4 \epsilon \left[ \left(\frac{\sigma}{r} \right)^{12} - \left(\frac{\sigma}{r} \right)^6 \right]+ \epsilon$ if $r < 2^{\frac{1}{6}}$, and zero otherwise \cite{weeks:5237}, with $\sigma$ the nominal particle diameter. We use $\epsilon = \kt$, but our results should be insensitive to the exact strength and form of the potential. $\Fp$ is the magnitude of the self-propulsion force which in the absence of interactions will move a particle with speed $\vp = D \beta \Fp$, $\hat{\boldsymbol{\nu}}_i = (\cos \theta_i, \sin \theta_i)$, and $\beta = \frac{1}{\kt}$. $D$ and $\Dr$ are translational and rotational diffusion constants, which in the low-Reynolds-number regime are related by $\Dr = \frac{3D}{\sigma^2}$. The $\eta$ are Gaussian white noise variables with $\left\langle\eta_i(t) \right\rangle= 0$ and $\left\langle\eta_i(t) \eta_j(t') \right\rangle= \delta_{ij} \delta(t-t')$. We non-dimensionalized the equations of motion using $\sigma$ and $\kt$ as basic units of length and energy, and $\tau = \frac{\sigma^2}{D}$ as the unit of time. Simulations employed the stochastic Runge-Kutta method \cite{PhysRevE.60.2381} with maximum timestep $2 \times 10^{-5} \tau$. Simulations mapping the phase diagram were run with $15{,}000$ particles until time $100 \tau$, while larger systems (up to $512{,}000$ particles) were used to explore kinetics and material properties. The simulation box was square with periodic boundaries, with its size chosen to achieve the desired density. The system is parametrized by two dimensionless values, the packing fraction $\phi$ and the P\'eclet number, which in our units is identical to the non-dimensionalized velocity ($\Pe = \vp \frac{\tau}{\sigma}$). In this work, we varied $\phi$ from near-zero to the hard-sphere close-packing value $\phi_{\mathrm{cp}} = \frac{\pi}{2 \sqrt{3}}$, and $\Pe$ from zero to $150$. \begin{figure}[tbp] \includegraphics[width=.51\columnwidth]{density_distribution.pdf} \raisebox{-1pt}{\includegraphics[width=.47\columnwidth]{dd.pdf}} \caption{(color online) Left: Phase densities as a function of P\'eclet number ($\Pe$) for a range of overall $\phi$. At low $\Pe$ the system is single-phase, while at increased $\Pe$ it phase-separates. The coexistence boundary is analogous to the binodal curve of an equilibrium fluid, with $\Pe$ acting as an attraction strength. Right: Observed density distributions for various P\'eclet numbers. In the single-phase region below $\Pe \approx 50$, $P(\phi)$ is peaked about the overall system density (here $\phi = 0.65$). It broadens and flattens as the critical point is approached, and becomes bimodal as the system phase separates.} \label{fig:phase_separation} \end{figure} \noindent \emph{Phase Separation}: We first show that our results are consistent with prior simulations \cite{PhysRevLett.108.235702} and confirm that this system, despite the absence of aligning interactions, shows the signature behaviors of an active fluid. In particular, the active spheres undergo nonequilibrium clustering (Fig.~(\ref{fig:summary})) similar to other model active systems \cite{Chate2008, PhysRevE.74.030904, C2SM06960A, PhysRevE.82.031904}. We establish that this clustering is indeed athermal phase separation by measuring the density in each phase at different parameter values (Fig.~(\ref{fig:phase_separation}a)). We observe a binodal envelope beyond which the system separates into two phases whose densities collapse onto a single coexistence curve which is a function of activity alone. The phase diagram is thus analogous to that of an equilibrium system of mutually attracting particles undergoing phase separation, with $\Pe$ (playing the role of an attraction strength) as the control parameter. This surprising result contradicts the expectation that increased activity will destabilize aggregates and suppress phase separation (as seen in \cite{Schwarz-Linek2012}) and indicates that the effects of activity cannot be described by an ``effective temperature'' in this system. Additionally, we identify a critical point at the apex of the bimodal (near $\Pe=50$, $\phi=0.7$). In the vicinity of this point, the system exhibits equilibrium-like critical phenomena which will be detailed in a future publication. \begin{figure}[tbp] \includegraphics[width=.49\linewidth]{cluster_phase.pdf} \includegraphics[width=.49\linewidth]{cluster_phase_theory.pdf} \caption{(color online) Left: Contour map of cluster fraction $\fc(\Pe, \phi)$ measured from simulations. The dashed curve marks the approximate location of the binodal. Right: Cluster fraction as predicted by our analytic theory (Eq. \ref{eq:cluster_fraction_theory}). These plots have been restricted to packing fractions that are low enough for the assumptions of our kinetic model to be valid, and for cluster identification to be unambiguous.} \label{fig:cluster_phase} \end{figure} \noindent \emph{The Phase-Separated Steady State}: To characterize the steady state, we measured the fraction of particles in the dense phase at time $100 \tau$ (Fig.~(\ref{fig:cluster_phase})). In contrast with recent work \cite{PhysRevLett.108.235702} which placed the phase transition boundary at a constant density, we observe that this cluster fraction is a nontrivial function of the system parameters $\fc(\Pe, \phi)$. To understand this relationship we developed a minimal model in which this function can be found analytically. Let us assume the steady state contains a macroscopic cluster which we take to be close-packed. Particles in the cluster are stationary in space but their $\theta_i$ continue to evolve diffusively. We treat the gas as homogeneous and isotropic, and assume that a particle colliding with the cluster surface is immediately absorbed. Within this model, we can write the rate of absorption of particles of orientation $\theta$ from the gas phase as $\kin(\theta) = \frac{1}{2 \pi} \rhog \vp \cos \theta$, where $\rhog$ is the gas number density. Integrating yields the total incoming flux per unit length: $\kin = \frac{\rhog \vp}{\pi}$. To estimate the rate of evaporation, note that a particle on the cluster surface will remain there so long as its self-propulsion direction remains ``below the horizon'', i.e., ${\bf\hat{n}}\cdot\boldsymbol{\hat{\nu}}<0$, where ${\bf\hat{n}}$ is normal to the surface. When its direction moves above the horizon, it immediately escapes and joins the gas. This rate can be calculated by solving the diffusion equation in angular space with absorbing boundaries (for clusters large enough to treat the interface as flat, at $\pm \frac{\pi}{2}$) and initial condition given by the distribution of incident particles: $\partial_t P(\theta, t) = \Dr \partial_\theta^2 P(\theta, t)$, with $P(\pm\frac{\pi}{2}, t) = 0$ and $P(\theta, 0) = \frac{1}{2} \cos \theta$. Further, the departure of a surface particle creates a hole through which subsurface particles (whose $\boldsymbol{\hat{\nu}}_i$ may point outwards) can escape. With $\kappa$ we denote the average total number of particles lost per escape event, which we treat as a fitting parameter. The total outgoing rate is then $\kout = \frac{\kappa \Dr}{\sigma}$. Equating $\kin$ and $\kout$ yields a steady-state condition for the gas density: $\rhog = \frac{\pi \kappa \Dr}{\sigma \vp}$. $\rhog$ can be eliminated in favor of $\fc$, yielding (in terms of our dimensionless parameters): \begin{equation} \label{eq:cluster_fraction_theory} \fc = \frac{4 \phi \Pe - 3 \pi^2 \kappa}{4 \phi \Pe - 6 \sqrt{3} \pi \kappa \phi} \end{equation} This function is plotted in Fig.~(\ref{fig:cluster_phase}) with $\kappa = 4.5$, in good accord with our simulation results. Further, the condition $\fc=0$ allows us to deduce a criterion for the onset of clustering. Restoring dimensional quantities, this condition gives $\phi \sigma \vp \sim \Dr$. Note that $\phi \sigma \vp$ is a collision frequency; thus the system begins to cluster at parameters for which the collision time becomes shorter than the rotational diffusion time. The mechanism we have presented here is purely kinetic and requires only an intuitive picture of local dynamics at the interface. An alternative view has been described by Tailleur and Cates \cite{Tailleur2008,2012arXiv1206.1805C} who subsume all interactions into a density-dependent propulsion velocity $v(\rho)$ which decreases with density as collisions become more frequent. From this they construct an effective free energy which shows an instability in the homogeneous phase if $v(\rho)$ falls quickly enough. In a sense our kinetic model represents an extreme case of this picture in which $v(\rho)$ contains a step function such that free particles are noninteracting, and particles in a cluster are completely trapped (see Fig.~\ref*{Supplement-fig:d_eff} in \cite{SupplementalInformation}). \begin{figure}[tbp] \raisebox{8pt}{\includegraphics[width=.445\columnwidth]{q6-defects.pdf}} \includegraphics[width=.54\columnwidth]{q6correlation.pdf} \caption{(color online) Left: Defect structures in a large cluster. Regions of high crystalline order (white) coexist with isolated and linear defects (dark). The color of each particle indicates its $|\q6|$. Inset shows pairs of 5/7 defects (red/blue). Right: Log-log plot of the correlation function $\langle \q6^*(\mathbf{r}) \q6(\mathbf{r}')\rangle$ for clusters at various P\'eclet numbers in systems with $N=128{,}000$, showing a transition from liquid-like exponential to hexatic-like power-law decay as activity is increased. For systems with $N=512{,}000$ (black dashed line), a crystal-like plateau is also observed.} \label{fig:q6} \end{figure} \noindent \emph{Structure of the Dense Phase}: Since the system is composed of monodisperse spheres, the dense phase is susceptible to crystallization \cite{PhysRevLett.108.168301}. As shown in Fig.~(\ref{fig:summary}) the static structure factor of the cluster interior shows a liquid-like isotropy at low $\Pe$, but develops strong sixfold symmetry as activity is increased. Further, the radial distribution function shows clear peaks at the sites of a hexagonal lattice (see Fig.~\ref*{Supplement-fig:radial_distribution} in \cite{SupplementalInformation}) which sharpen and increase in number as $\Pe$ is raised. We also measured the bond-orientational order parameter $\q6(i) = \frac{1}{|\mathcal{N}(i)|} \sum_{j \in \mathcal{N}(i)} e^{i 6 \theta_{ij}}$, where $\mathcal{N}(i)$ runs over the neighbors of particle $i$ (defined as being closer than a threshold distance), and $\theta_{ij}$ is the angle between the $i$-$j$ bond and an arbitrary axis (Fig.~(\ref{fig:q6})). We find a structure characterized by large regions of high order with embedded defects that are predominantly 5-7 pairs (Fig.~(\ref{fig:q6}a) inset and \ref*{Supplement-fig:defects-movie} in \cite{SupplementalInformation}). Next, we examined the correlation function $\left\langle\q6^*(\boldsymbol{r}) \q6(\boldsymbol{r}') \right\rangle$ (Fig.~\ref{fig:q6}) which exhibits a liquid-like exponential decay for systems of low activity, while at higher activity the decay slows to a power law which is indicative of a hexatic \cite{2002dgcm.bookN}. A further transition to a crystal-like plateau is observable in larger systems (see Fig. (\ref{fig:q6}) and \ref*{Supplement-fig:q6correlation-size} and \ref*{Supplement-fig:defect-density} in \cite{SupplementalInformation}). In all cases, this material is unique in that it is held together by active forces alone, and that the arrest of motion is due to frustration. In this sense it is similar to amorphous materials such as granular packs as reflected by the highly heterogeneous stress distribution (Fig.~(\ref{fig:summary})) \cite{doi:10.1021/jp809768y}. \noindent \emph{Dynamics in the Dense Phase}: Within the active solid material, self-propulsion forces continuously evolve by rotational diffusion, breaking local force balance and leading to defect formation and migration (see \ref*{Supplement-fig:defects-movie} in \cite{SupplementalInformation}). A compelling way to view the motion produced by this athermal process is a simulated FRAP experiment \cite{blaaderen:4591}, in which particles within a contiguous region are tagged, making subsequent mingling of tagged and untagged particles visible (see \ref*{Supplement-fig:frap-2.avi} in \cite{SupplementalInformation}). To quantify this behavior, we measured the mean square displacement (MSD) of particles in the cluster interior. As shown in Fig.~(\ref{fig:summary}), we observe subdiffusive motion on short timescales, followed by a superdiffusive regime, returning to diffusive motion on long timescales. The exponents of the subdiffusive and superdiffusive motion ($\frac{1}{2}$ and $\frac{3}{2}$, respectively) are well-conserved across a wide range of propulsion strengths. Note that an isolated self-propelled particle will exhibit diffusive, ballistic and diffusive behavior on time scales $t < \frac{4 D}{\vp^2}$, $\frac{4 D}{\vp^2} < t < \frac{1}{\Dr}$ and $t > \frac{1}{\Dr}$ respectively (see Fig.~\ref*{Supplement-fig:msd} in \cite{SupplementalInformation}). These dynamical regimes are modified by the active solid environment; in particular, the ballistic regime is modulated by ``sticking'' events as the particle is localized in crystal domains, resulting in the observed L\'evy-flight-like behavior\cite{klafter:33, doi:10.1080/00018737800101474}. \begin{figure}[tbp] \includegraphics[width=0.495\linewidth]{nucleation_growth.pdf} \includegraphics[width=0.485\linewidth]{coarsening.pdf} \caption{(color online) Examples of phase separation kinetics. Left: A system with $\Pe=100$, $\phi=0.45$ in which a delayed nucleation event leads quickly to steady-state. For shallowly-quenched systems, the nucleation time can be long enough that artificial seeding is needed to make nucleation computationally accessible. Right: A system with $\Pe=80$, $\phi=0.6$ where spinodal decomposition leads to a coarsening regime which slowly evolves towards steady-state. Inset shows mean cluster size scaling approximately as $t^\frac{1}{2}$. (see \ref*{Supplement-fig:coarsening-movie} and \ref*{Supplement-fig:coarsening} in \cite{SupplementalInformation}).} \label{fig:nucleation_growth_coarsening} \end{figure} \noindent \emph{Kinetics of Phase Separation}: Despite the athermal origins of phase separation in this system, simulations quenched to parameters within the binodal experience familiar phase separation kinetics (Fig.~(\ref{fig:nucleation_growth_coarsening})). Systems quenched close to the binodal exhibit a nucleation delay which can be long enough that artificial seeding is necessary for phase separation to be computationally accessible. Systems quenched more deeply undergo spinodal decomposition, leading to a coarsening regime in which the mean cluster size scales surprisingly as $t^\frac{1}{2}$, with a corresponding length scale $\mathcal{L}(t) \sim t^\frac{1}{4}$ (Fig.~(\ref{fig:nucleation_growth_coarsening}) inset, also see \ref*{Supplement-fig:coarsening} in \cite{SupplementalInformation}). This differs from the standard 2D coarsening exponents, but matches recent simulation results for the Vicsek model and related active systems \cite{PhysRevLett.108.238001}. This result should be viewed as preliminary due to the limited range of our data, but nevertheless this unexpected similarity between the coarsening of point-particles with polar alignment and that of spheres with no alignment suggests a deep relationship between these very different types of systems. Future work is needed to uncover the origins of these scaling exponents and their implications for universality in active fluids. \noindent \emph{Summary}: A fluid of self-propelled colloidal spheres exhibits the athermal phase separation that is intrinsic to active fluids and is a primary mechanism leading to emergent structures in diverse systems \cite{Gopinath2011,Cates2010}. We have shown that the physics underlying this phase behavior can be understood in terms of microscopic parameters. From a practical perspective, our simulations show that the active solid dense phase exhibits a combination of structural and transport properties not achievable in a traditional passive material. Further development of experimental realizations of this system (e.g. Ref. \cite{2012arXiv1202.6264T}) will advance the development of materials whose phase behavior, rheology, and transport properties can be precisely controlled by activity level. {\bf Acknowledgments:} This work was supported by NSF-MRSEC-0820492 (GSR, MFH, AB), as well as NSF-DMR-1149266 and NSF-1066293 and the hospitality of the Aspen Center for Physics (AB). Computational support was provided by the Brandeis HPC. \bibliographystyle{apsrev4-1} \bibliography{paper} \end{document} }

75 — 1207.2447

\caption{\label{fig:NT_YoungModulus}Calculated radial Young's modulus of nanotubes of different sizes. {\color{red} }}

76 — 1207.2686

\caption{\label{pic:dilution}(Color online) a) The central part of the detector system. The CsI(Tl) crystals (blue and green) are read out via wavelength shifters and photodiodes or photomultipliers, BaF$_2$ crystals (yellow) in forward direction with photomultipliers. b) The missing mass distribution, and (c) the $\pi^0$ angular distribution for reaction (1) for an incident photon energy of $E_\gamma = 1000\pm25$~MeV; butanol ({\scriptsize$\square$}), hydrogen ({\color{blue}$\triangle$}), carbon ({\color{red}$\circ$}), and the sum of hydrogen and carbon data ({\color{green}$\ast$}). From these distributions, the dilution factor (d) is determined.}

\caption{\label{pic:Sigma}(Color online) a) A typical $\phi_\pi$-distribution with a fit using eq. (2). b,c) The beam asymmetry $\Sigma$ as a function of $\cos\theta_{\pi}$ for $E_\gamma =800$\,MeV and for$E_\gamma =1100$\,MeV. Black dots show our data, red squares GRAAL data\cite{Bartalini:2005wx}. The curves represent predictions from different partial wave analyses. Solid (black) curve: BnGa \cite{Anisovich:2012}; dashed (red): SAID \cite{Dugger:2009pn}; long-dashed (black): BnGa with $E_{0^+}$ and $E_{2^-}$ amplitudes from SAID; dotted (blue): MAID \cite{Drechsel:1998hk}; dashed-dotted (black): BnGa with $E_{0^+}$ and $E_{2^-}$ amplitudes from MAID. Gray area shows the systematic error due to interactions on nuclei and uncertainty in the photon polarization.}

\caption{\label{pic:results}(Color online) The polarization observable $G$ as a function of $\cos\theta_{\pi}$ from $E_\gamma =800$\,MeV up to$E_\gamma =1100$\,MeV. Systematic errors are shown in gray bars. Curves: see fig.\protect\ref{pic:Sigma}.}

77 — 1207.2711

\caption{ An example network, which is drawn according to the uniform clustering model, is shown in the upper right portion of the figure. The reference receiver (indicated by the {\color{red} $\star $}) is placed at the origin, and the corresponding reference transmitter is located to its right (indicated by the {\color{red} $\bullet $} to the right of the receiver). $% M=28$ interferers (indicated by the {\color{blue} $\bullet $}) are placed at random within the dashed circle. An exclusion zone surrounds each mobile (indicated by solid circles surrounding the mobiles). The figure also shows the outage probability as a function of SNR $\Gamma $, conditioned on the pictured network topology. Performance is shown for three fading models without spreading or shadowing. Analytical expressions are plotted by lines while dots represent simulation results (with one million trials per point).}

78 — 1207.2796

\caption{Computed IR Spectrum of ethylene (a) and projection of the energy gradients of singly oxidized ethylene (b) and its deuterated isotopomers in the 2000-3500 cm$^{-1}$ CD and CH stretching region. Color key: {\color{mma1}--} $C_{2}H_{4}$, {\color{mma2}--} $DHC_{2}=C_{2}H_{2}$, {\color{mma3}--} $D_{2}C_{2}=C_{2}H_{2}$}

79 — 1207.2825

\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 while 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. \vspace{-0.65cm} }

80 — 1207.7349

\caption{\coloronline Schematic of apparatus showing (a) two walls bi-axially compressing an array of disk-shaped particles composed of an outer subsystem (black, high $\mu$) and an inner subsystem (red, low $\mu$) and (b) reflective photoelasticity on air-floated particles. Light shines from green LEDs through a linear polarizer (P) a wavelength-matched quarter wave plate (Q) before entering the photoelastic material. A mirrored surface on the bottom of each particle reflects light back through the particle. A second quarter-wave plate and linear polarizer are mounted on the camera to resolve the photoelasticity. Three images of each configuration are recorded: (c) unpolarized white light for locating particle positions, (d) polarized green light showing isochromatic fringes for calculating contact forces and (e) an ultraviolet light for identifying the low-$\mu$ particles.}

\caption{\coloronline (a) Volume histograms, ${\cal P}(V)$, for $\Phi = 0.776$ ($\blacktriangle$), $0.784$ ($\blacksquare$), and $0.802$ ($\bullet$) with $m=48$. (b) A semi-logarithmic plot of the ratio each histogram with respect to the $\Phi = 0.784$ distribution, i.e. ${\cal P}_i(V)/{\cal P}_{i=2}(V)$. (c) The inverse compactivity given by \eref{eqn:volume} plotted as a function of the inverse volume fraction where $\mu_B$ are shown as black \textbullet ~and $\mu_S$ are red $\blacklozenge$. Large/small symbols denote jammed/un-jammed configurations, respectively. Errorbars shown are uncertainties in ${\cal P}(V)$ and propagated through the calculation. The inverse compactivity given by the FDT method (\eref{eqn:fdt}), is shown with the solid line for comparison. (d) The ratio of number of jammed/un-jammed configurations recorded at each $\Phi$. }

\caption{\coloronline (a) Distribution of $\sigma_p$ where $m=8$ and $\GammaN = 0.0007$ ($\blacktriangledown$), $0.0010$ ($\blacksquare$), $0.0015$ ($\bullet$), and $0.0024$~Nm ($\blacktriangle$). A semi-logarithmic plot of the (b) ratio ${\cal P}_i(\sigma_p)/{\cal P}_j(\sigma_p)$ where the reference system $j$ is $\GammaN = 0.0015$~Nm. The pressure angoricity $A_P$ and shear angoricity $A_\tau$ are shown as a function of $\GammaN$ where the results using overlapping histograms for $\mu_B$ and $\mu_S$ are shown as black $\circ$ and are red $\diamondsuit$, respectively. The solid line is the angoricity calculated using FDT. The gray dashed lines provide a visual reference of the slopes $0.15$ and $0.45$, respectively. Inset: The scaled variance $\langle \delta\sigma_p^2 \rangle$ of the ${\cal P}_j(\sigma_{p})$ distribution, as a function of the cluster size $m$.}

81 — 1208.2239

\caption{An example of load balancing where $n=2$, $p_1=0.55$, $p_2=0.45$, $k=3$, and $N_\mathrm{w}=4$. The second to last line denotes the load of each vertex and the last line denotes the cumulative load. The 4 processors attempt to take $25\%$ of the total load each. Consequently processor number {\color{Blue}0}, {\color{Green}1}, {\color{RoyalPurple}2}, and {\color{Red}3} takes ownership to the {\color{Blue}blue}, {\color{Green}green}, {\color{RoyalPurple}purple}, and {\color{Red}red} vertices, respectively.}

82 — 1208.2421

\caption{Simulation results comparing the frequency response of four planar loop designs: On-chip CPS with no transitions (\chain); On-chip CPS with PCB balun (\dashed); On-chip CPW (\dotted); On-chip balun (\full). (a)~S$_{11}$ parameter. (b)~Magnitude of the perpendicular component of the magnetic field at the donor site. (c)~Absolute electric field at the charge detector, in this case the SET island.}

\caption{(a)~Comparison of S$_{11}$ parameters for different bond wire lengths and types: 200~$\mu$m bond wire (\full); 500~$\mu$m bond wire (\dashed); 150~$\mu$m bond wire (\dotted); 200~$\mu$m ribbon bond 70~$\mu$m wide by 10~$\mu$m thick (\chain). (b)~Comparison between: the standard model with the qubit located outside the loop (\full); model with the transmission line elevated with calixarene, with the qubit located inside the loop (\dashed); standard model with a thin film of calixarene surrounding the transmission line and gates (\chain). (c)~Electric and magnetic fields obtained from the model with (\full) and without (\dashed) the bond wires bridging the ground plates of the on-chip CPW. The coplanar structure under study on all subfigures is the on-chip balun shown in \fref{fig:ChipBalun}. The device and probe locations with respect to the substrate are identical in all models.}

83 — 1208.2518

\caption{\label{fig_toy}(left)~A simple Java class and the corresponding part of class dependency network. Direction of links is (mostly) just the opposite to the flow of information. (right)~Class dependency network of \texttt{java} ({\color{red}circles}) and \texttt{javax} ({\color{gray}triangles}) namespaces of Java language.}

\caption{\label{fig_jung}A random graph, \jng network, \jng \&\clt network and \jng \&\jav network. Average distance between the nodes $l$ equals $3.88$, $4.19$, $5.37$ and $2.18$. Node symbols correspond to clustering $D$~\cite{SV05} that ranges between $0$ ({\color{gray}triangles}) and $1$ ({\color{red}circles}).}

\caption{\label{fig_abst}(left)~\jng network where node symbols represent high-level packages of JUNG framework: \texttt{visualization} ({\color{red}circles}), \texttt{io} ({\color{blue}triangles}), \texttt{graph} ({\color{green}squares}) and \texttt{algorithms} ({\color{gray}diamonds}). (right)~Hierarchy of structural modules revealed with the algorithm in~\cite{SB12f}.}

84 — 1208.3450

\caption{{\color{red}{(Color online.) }} Examples of relaxed structures of trans-PDT molecules bonded to gold clusters in the top geometry.\cite{Macmolplt} (a): A PDT molecule with no hydrogen atom attached to either sulfur atom. (b): A PDT molecule with a hydrogen atom attached to one of the sulfur atoms. (c): A PDT molecule with a hydrogen atom attached to each of the two sulfur atoms. (d): Two PDT molecules bridging the gold clusters with no hydrogen atom attached to any sulfur atom. (e): Two PDT molecules bridging the gold clusters with a hydrogen atom attached to one of the sulfur atoms of each molecule. The sulfur atom of the molecule on the right has not chemisorbed to the lower gold cluster while each of the other three sulfur atoms in the system has chemisorbed to a gold cluster. (f): Two PDT molecules bridging the gold clusters with a hydrogen atom attached to each of the two sulfur atoms of each molecule. Carbon, hydrogen, sulfur and gold atoms are black, blue, yellow and amber, respectively. }

\caption{{\color{red}{(Color online.) }} Calculated low bias elastic conductances, and mode I phonon energies and IETS intensities for the relaxed structures of one and two molecules with different numbers of thiol hydrogen atoms bridging gold atomic clusters. The numerical values labelled (a) -- (f) are for the molecular junctions shown in Fig. \ref{Fig1} (a) -- (f), respectively. The experimentally measured low bias conductance values of Hihath \emph{et al.} for single-molecule PDT junctions were $\sim 0.006 g_{0}$ $\pm 0.002$. \cite{HihathArroyoRubio-BollingerTao08} $g_0 = 2e^2/h$. }

\caption{{\color{red}{(Color online.) }} \subref{Fig2(a)} Calculated vibrational modes in the energy range from $25$ to $60$~meV for PDT bridging gold nano-clusters with hydrogen atoms bound to one or both sulfur atom(s), the sulfur atoms being bonded to gold in top-site geometries. These structures were relaxed with no constraints imposed on their geometries. Carbon, hydrogen, sulfur and gold atoms are black, blue, yellow and amber, respectively.\cite{Macmolplt} Arrows show un-normalized atomic displacements; the heavier arrows indicate the motion of the sulfur atoms. %%% \subref{Fig2(b)} Calculated IETS intensities (colored) vs. calculated vibrational mode energies for PDT molecules linking pairs of gold clusters with between 12 and 14 Au atoms in each cluster. Results are shown for both sulfur atoms bonding to the gold in top-site geometries and having no hydrogen atom attached to either sulfur atom, having a hydrogen atom attached to one sulfur atom only, and having a hydrogen atom attached to each of the sulfur atoms. % The experimental IETS phonon mode histogram of Hihath \emph{et al.}\cite{HihathArroyoRubio-BollingerTao08} is shown in (darker, lighter) grey for (positive, negative) bias voltages. The energy ranges in which modes of types I, II, III, and IV in \subref{Fig2(a)} occur are indicated by arrows. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% }

\caption{{\color{red}{(Color online.) }} Calculated inelastic tunneling spectra for one and two PDT molecules with different numbers of thiol hydrogen atoms bridging pairs of 14 atom gold clusters in the fully relaxed geometries depicted in Fig. \ref{Fig1}. Labels (a) -- (f) denote the structures (a) -- (f) of Fig. \ref{Fig1}. The atomic motions in modes I -- IV are similar to those shown for modes I -- IV of the single molecules in Fig. 2(a). }

\caption{{\color{red}{(Color online.) }} \subref{Fig4(a)} I to IV: Calculated geometries of PDT molecular junctions with a H atom bound to one S atom of the molecule. C, H, S and Au atoms are black, grey, yellow and amber, respectively. In the relaxations the separation of the atoms of the two Au clusters furthest from the molecule was held fixed. Au--S distances are shown. V: Molecular junction relaxed without constraints and with no H atom attached to either S atom. \subref{Fig4(b)} Theoretical inelastic tunneling spectra of the PDT wires, some of which are shown in \subref{Fig4(a)}, color coded according to the Au--S distance for the S atom with an attached H atom. The calculated low bias elastic conductance of each structure is shown in the legend. Chartreuse arrows indicate mode IV. Black arrow indicates the energy of mode IV if there is no Au cluster near the S--H group. Pink areas indicate energy ranges in which modes I, II and III occur. The experimental phonon mode histogram (grey) of Hihath \emph{et al.}\cite{HihathArroyoRubio-BollingerTao08} is shown as a reference. \subref{Fig4(c)} Au-S and thiol H vibrational modes are plotted vs. Au-S distance; Au-H distances are also shown. Mode IV points are circled in black. The solid red curve is a fit of the mode IV frequencies to an inverse cube dependence on the Au--S distance; see text. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% }

85 — 1208.3806

\caption{A Markov chain describing \blue{transitions in the Markov state of a receiver $r$.}}

\caption{\blue{Probability of coefficient-based delivery as a function of Markov state, for baseline transmission schemes with addition rate $\lambda=0.7$.} Probabilities are normalised over coefficient-based \red{deliverable} timeslots. The field size $M=R$ the number of receivers, and $\mu=0.8$.}

\caption{The average delivery delay of \blue{coding} schemes A, B and \blue{RLNC under the baseline rate control scheme with addition rate $\lambda$}. Zero and leader state \blue{delivery delays} are included for comparison. $R=4,\mu=0.8$.}

\caption{For the baseline rate control scheme with coding scheme B, the \blue{probability that $N$} packets are coded \blue{into a} sender transmission. $R=4,8$, $\lambda=0.7$, $\mu=0.8$. Note that, \blue{if the transmission queue is empty, no packets will be} coded.}

\caption{The probability from \eqref{zerotozero} of a receiver taking $\leq T$ time steps to return to the zero state \red{under the baseline rate control scheme}.}

\caption{The \blue{zero state delivery delay of the baseline rate control scheme}, as a function of the addition rate $\lambda$. \blue{The \grn{delay estimates} of} \eqref{zdeccalc} (dotted lines) \blue{are compared against} simulation (solid lines).}

\caption{\blue{Under the baseline rate control scheme with addition rate $\lambda$,} the proportion of time the leader is in each state, under the independent receiver model \eqref{leadindep} and in practice. $\lambda=0.7,\mu=0.8$.}

\caption{\blue{Simulated zero state and leader state \blue{delivery delays under the baseline rate control scheme with addition rate $\lambda$}, for different numbers of receivers $R$.} $\mu=0.8$.}

\caption{\blue{Throughput-delay \grn{performance} for the coding scheme B transmission schemes. \bII{From left to right, the parameters for each transmission scheme are:} Baseline: $\lambda=0.5,...,0.75$. Delay threshold: $T_D=2,...,100$. \bII{Dynamic: $f=2,...,500$.} For all transmission schemes $R=4$ and $\mu=0.8$.}}

86 — 1208.4349

\caption{ Nucleon masses (GeV) using LMA, AMA and from data from a high statistics study \cite{Yamazaki:2009zq}. {\red See Table~\ref{tab:cost} for costs. Gaussian (gauss) and point (pt) sinks.}}

87 — 1208.4798

\caption{Variation of the height of the center of mass of the fluid (liquid plus vapor) as a function of time. Initial velocity 1.0 (\red{$*$}), and 2.0 $\sigma/\tau$ (\blue{$\times$}) on a non-wetting surface and velocity 1.0 $\sigma/\tau$ (\green{$\circ$}) on a wetting surface.}

\caption{(a) $p$-$V$ diagram for dimers; points (blue) are simulation results and the line (red) is a fit. (b) Density profile produced by an oscillating wall: {\red{$\circ$}} oscillating wall, {\blue{$\times$}} stationary wall.}

88 — 1208.5169

\caption{(color online) Orientational polarization~\eqref{eq:aggregatedP} for~$\alpha^*=2$ ({\color{blue}{\bf -}}) and the approximation~\eqref{eq:pdf_to_L} ({\color{red}{\bf -\,-}}). \label{fig:fAndL} }

\caption{(color online) Comparison of the predicted dielectric constant~\eqref{eq:thePrediction_eps}, with experimental data as function of ionic concentration~$c$ for various salts. The theoretical prediction was calculated using~$\beta$ as a fitting parameter. A: Data for NaCl salt from~\cite{Hasted1948}, compared to~\eqref{eq:thePrediction_eps} with~$\beta=42.07$. B: Data from~\cite{Wei92}, where fit for RbCl and CsCl salt ({\color{red} $--$}) is obtained with~$\beta=67.20$ and for LiCl ({\color{blue}$-$}) with~$\beta=53.22$. C: Figure 2(b) from~\cite{Andelman2012} where fit for KF ({\color{blue} $--$}) is obtained with~$\beta=75$ and with~$\beta=60$ for KCl ({\color{oliveGreen} $.-$}). Solid black curve is the prediction obtained using the field-theory approach~\cite{Andelman2012}. \label{fig:dielectricDeptFit} }

\caption{(color online) Comparison of the predicted dielectric constant~\eqref{eq:thePrediction_eps}, with experimental data from~\cite{Buchner99} as a function of ionic concentration~$c$ for NaCl at various temperatures. Data for~$T=5^\circ$ fit with~$\alpha=13.7$ and~$\beta=70.25$ ({\color{red}{\bf -\,-}}), for~$T=20^\circ$ with~$\alpha=12$ and~$\beta=52.94$ ({\color{blue}{\bf -}}), for~$T=25^\circ$ with~$\alpha=11.5$ and~$\beta=47.91$ ({\color{oliveGreen}{\bf .-}}) and for~$T=35^\circ$ with~$\alpha=10.7$ and~$\beta=40.12$ ({\color{magenta}{\bf $\times$-}}). \label{fig:NaCl_temp} }

89 — 1208.5899

\caption{Energy spectra of the detected X-ray photons. The three sets of data \opentriangle, \opensquare and \opencircle correspond to the first, second and third x-ray burst respectively.}

90 — 1208.6415

\caption{Three-body scaling plot for various systems as marked in the legend. The contribution to the mean square radius from the three-body radial coordinate is measured against the scaling radius from equation (\protect\ref{eq:rho2}) and displayed versus the scaled binding energy. The arrows indicate where two-body subsystems become bound and the dashed line (\dashed) indicates where Efimov states can appear. See the text for details. Courtesy E. Garrido. \label{fig:scale3}}

91 — 1209.0218

\caption{Oscillation frequencies detected for 18 Sco along with their radial order $n$ and degree $l$. The quantities given here are the median of the marginal posterior distributions for each eigenfrequency considered in our model. The corresponding credible intervals are given alongside. We use some flags to signal estimates requiring additional caution. (\textasteriskcentered) indicate eigenfrequencies with clear multimodal marginal distributions, (\dag) pulsation modes potentially unresolved, (\textdaggerdbl) pulsation modes that have not passed a hypothesis testing when using the heights estimated with the MAP approach in Sect.~\ref{sect:data_map} and (?) the mode that has not been identified by the MAP approach.}

92 — 1209.1226

\caption{(Color online) (a) Hilbert-Schmidt norm (\ref{eq:schmidt-hilbert}) vs. feedback parameter $\Theta$ for single qubit, $\Gamma'/\Gamma =$~2.8 and $\vert\varphi\vert =$~0.4234$\pi$, $g_+$ (red, dashed) and $g_-$ (blue, solid). The shape of the curves around $g_k=$~0 provides the stabilization sensitivity on $\Theta$. (b) Feedback angle $\varphi$ and (c) Feedback parameter $\Theta$ [see (\ref{eq:singlefeed})] vs. coupling strength $\Gamma'/\Gamma$ required for perfect feedback stabilization $g_k=$~0. Different curves correspond to $\varepsilon=$~0 (blue, solid), $\varepsilon=$~$-$0.1 (orange, solid), $\varepsilon=$~0.1 (orange, dashed), $\varepsilon=$~$-$0.2 (green, solid), $\varepsilon=$~0.2 (green, dashed). Symbol \textcolor{red}{$\bullet$} indicates parameter set of red(dashed) and blue(solid) curves in (a). For finite dissipation perfect stabilization fails, see thin black curves in (a) $1/(\tau\Gamma) =$~0.02, 0.1. Parameters: $U=$~0, $T_C/\Gamma=$~1. }

93 — 1209.1529

\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'. Lane changes, whether in the incoming street segments or in the intersection square, are forbidden.}

\caption{\small Zoom on a region of fig.\,\ref{fig_alpha640_0.09}. To emphasize their difference the northbound particles (\orangew{$\blacktriangle$}) have been indicated by larger symbols than the eastbound particles (\bluew{$\blacktriangleright$}). The black solid lines are at an angle of $45^\circ$. The nonzero distances $\Delta_1,\Delta_2,$ and $\Delta_3$ show that locally $\Delta\theta(\br)>0$. }

\caption{\small Mechanism causing the deviation $\Delta\theta$ of a stripe of northbound particles (\orangew{$\blacktriangle$}) in the upper triangular region. }

94 — 1209.1627

\caption{(Color online). The shift in $\nu_{\rm R}(\Omega)$ for the $D=0.26$\,mm slab due to the flow-induced\Fred transition for both directions of rotation. The lines (solid blue: $\mathbf{\Omega}\parallel -\bl$, dashed red: $\mathbf{\Omega}\parallel\bl$) are calculated using Eqn.\,\ref{freqshift} for the two possible types of oriented textures (top: $\bl=-\Omhat$ as on Fig.\,\ref{fig2}(a); bottom (shifted by -1.5\,mHz for clarity):$\bl=\Omhat$ as on Fig.\,\ref{fig2}(b)). They are compared with the typical experimental data, $\mathbf{\Omega}\parallel\pm\mathbf{\Omega_0}$ (same data points are also repeated shifted by -1.5\,mHz).$T=0.9 T_{\rm c}$ and $C=-18$\,mHz.}

95 — 1209.1724

\caption{Permittivity of silver as a function of temperature for\textcolor{red}{{} }$\lambda_{0}=\unit[1550]{nm}$ (blue line) and $\lambda_{0}=\unit[650]{nm}$ (red line). We see that an increase of about 100 is possible in the first case, while an increase of about 10 is possible for optical frequencies.\label{fig:Temperature}}

96 — 1209.1993

\caption{\red{(a)}}

\caption{\red{(b)}}

\caption{\red{(c)}}

97 — 1209.2784

\caption{\small Maximum RMSE (Left) and normalized mean RMSE (Right) versus task-specific parameter bound $\tau_1$, for shared parameter bound $\tau_0$ fixed. In each figure, Left section is $\tau_0$ is 0.2 and Right section is $\tau_0 = 0.6$. Solid red {\scriptsize \textcolor{red}{$\blacklozenge$}} is $\ell_1$, solid blue \textcolor{blue}{$\bullet$} is minimax, dashed green \textcolor{OliveGreen}{$\blacktriangle$} is $(0.1 T)$-minimax, dashed black $\blacktriangledown$ is $(0.2 T)$-minimax. The results for $\ell_2$ MTL were visually identical to $\ell_1$ MTL and hence were not plotted. }

\caption{\small MTL (Top) and LTL (Bottom). Maximum $\ell_2$ risk (Left) and Mean $\ell_2$ risk (Right) vs bound on $\|W\|_{\mathrm{tr}}$. LTL used 10-fold cross-validation (10\% of tasks left out in each fold). Solid red {\scriptsize \textcolor{red}{$\blacklozenge$}} is $\ell_1$, solid blue \textcolor{blue}{$\bullet$} is minimax, dashed green \textcolor{OliveGreen}{$\blacktriangle$} is $(0.1 T)$-minimax, dashed black $\blacktriangledown$ is $(0.2 T)$-minimax, solid gold {\scriptsize \textcolor{BurntOrange}{$\blacksquare$}} is $\ell_2$. }

98 — 1209.3897

\caption{\textbf{(a)} Mean-square displacement of a Brownian particle in a simple shear along the $y$ direction for $a=1.00$ ({\color{myblue}{\large $\bullet$}}) and $5.00$ ({\color{mypurple} $\blacksquare$}). The corresponding mean-square displacement along the $x$-direction is shown for $a=1.00$ ({\Large \color{myblue} $\circ$}) only. The solid lines are the plots of \fref{eq:msdy} for the respective values of $a$ and the dot-dashed line is the plot of $\la x^2(t)\ra= 2 D t$. \emph{\textbf{Inset:}} Plot of $\la x(t) y(t)\ra$ for $a=1.00$ ({\color{myblue}{ $\vartriangle$}}) and $5.00$ ({\color{mypurple} $\triangledown$}). The dashed line are the plots of \fref{eq:xyt}. \textbf{(b)} Persistence probability of a Brownian particle in a simple shear for ({\color{myblue}{\large $\bullet$}}), $5.00$ ({\color{mypurple} $\blacksquare$}) and $10.0$ ({\color{myokker} $\blacktriangle$}). The dashed line is plot of $t^{-1/2}$ corresponding to that of a free Brownian particle, while the dot dashed line is a plot of $t^{-1/4}$. }

\caption{Mean-square displacement of a Brownian particle in a simple shear along the $y$ direction for $a=1.00$ ({\color{myblue}{\large $\bullet$}}), $2.50$ ({\color{mypurple} $\blacksquare$}) and $5.00$ ({\color{myokker} $\blacktriangle$}). The corresponding mean-square displacement along the $x$-direction is shown for $a=1.00$ ({\Large \color{myblue} $\circ$}) only. The solid lines are the plots of \fref{eq:msd_hc_y} for the respective values of $a$ and the dot-dashed line is the plot of $\la x^2(t)\ra= 2 D t$. \emph{\textbf{Inset:}} Plot of $\la x(t) y(t)\ra$ for $a=2.50$ ({\color{mypurple}{ $\vartriangle$}}) and $5.00$ ({\color{myokker} $\triangledown$}). The dashed line are the plots of \fref{eq:hc_xy_t}.}

\caption{Persistence probability of a semi-flexible polymer in a shear flow for values of $a=0.1$ ({\color{myblue}{\large $\bullet$}}) and $0.50$ ({\color{mypurple} $\square$}). The solid black line is a plot of $t^{-\theta}$, with $\theta=0.29695$, the prediction of IIA. A power law fit to the asymptotic of the data yields $\theta=0.306236$ for $a=0.1$ and $\theta=0.297205$ for $a=0.5$. }

99 — 1209.5195

\caption{(a) Time-averaged electric field (obtained from the simulation) near the apex of a tungsten tip ($R = 30\,\mathrm{nm}$, $\lambda = 800\,\mathrm{nm}$) in a plane spanned by the tip axis and the wave vector $\textbf{k}$ of the laser. The tip shape is indicated as a gray line. The white line at the bottom displays the near-field along $z = 0$ with the $1/e$ decay length $L$. The field rises from 1.2 to 3.4 over a distance of $29\,\mathrm{nm}$, where $1$ indicates the field strength in the bare laser focus without a tip. Note that the near-field is not symmetric with respect to the tip axis. This asymmetry is more prominent for larger tip radii~\cite{Yanagisawa2010}. (b) Illustration of electron rescattering: electrons are emitted in the optical near-field of a metal nanotip. A fraction of the emitted electrons is driven back to the tip surface, where they can scatter elastically. The kinetic energy gained during the rescattering process depends sensitively on the electric field near the tip surface. Thus the strength of the optical near-field is mapped to the kinetic energy of the emitted electrons. (c) Typical energy distribution of electrons emitted in the forward direction. The high-energy plateau ($\sim 5$ to $10\,\mathrm{eV}$) arises due to rescattering. Its cut-off is related to the local electric field amplitude at the metal. We obtain it from the intersection of two exponential fit functions (red lines). (d) Decay length $L$ as function of tip radius $R$ for tungsten tips ({\color{navblue} $\filledmedsquare$}) and gold tips ({\color{red} $\medbullet$}), deduced from simulations. The lines show linear fits: $L = (0.90 \pm 0.03) R$ for tungsten and $L = (0.82 \pm 0.04) R$ for gold. As the shape of the near-field mainly depends on the tip geometry, other materials behave very similarly.}

\caption{Experimental results for the field enhancement factor of tungsten tips ({\color{navblue} $\medbullet$}) and gold tips ({\color{red} $\filledmedsquare$}) as a function of the tip radius. The uncertainty in $\xi$ represents an estimated systematic error due to the uncertainty in laser intensity. The lines are simulation results for $\lambda = 800\,\mathrm{nm}$ (W: solid blue line, Au: dashed red line, Ag: dash-dotted black line). The dielectric functions of the metals are taken from experimental data~\cite{Lide2004} (see Fig.~\ref{epsilon}(b)). For technical reasons related to mock surface plasmon reflection, we simulate gold and silver tips with a smaller opening angle than tungsten tips (W: $5\degree$, Au, Ag: $0\degree$). Simulations of tips with different angles show that this should not alter the results by more than $5\%$.}

\caption{Determination of tip radii by field ion microscopy and scanning electron microscope imaging. (a) Field ion microscope (FIM) image for the tungsten tip with a smaller radius of curvature. Counting the number of rings $n$ in the image between two crystallographic poles (in this case (110) and (211)) gives an estimate of the tip radius of curvature $R$ (here $n = 8\pm 1$, or $R = (13.4\pm 1.7)$\,nm). (b) The same for the tip with a slightly larger radius (here$n = 9\pm 1$, or $R = (15.0\pm 1.7)$\,nm). (c) Ball model of a tungsten tip in (310) orientation. The color map indicates the deviation of the surface atoms from an ideal hemisphere. Protruding atoms are visible as bright spots in FIM images. The ring counting method is applied between the (110) and (211) poles (dotted line). (d) Scanning electron microscope image of a gold tip ($R = (46 \pm 3)$\,nm).}

100 — 1209.5743

\caption{\label{FIG:phases} \co (Upper panel)~The energy per site versus the coupling strength. The solid lines are best energies from the trial wave function optimization. From bottom to top, system sizes $L = 4, 6, 8, 10, 12, 16, 20, 24, 28, 32$ are shown. The thick black line, providing an upper envelope to the curves, is the extrapolation to $L=\infty$. The ground state energy of the one-dimensional Heisenberg chain is shown for comparison, as are the energies of the NNRVB state and its $45^{\circ}$-rotated, next-nearest-{\neighbour} {\analogue}, the NNNRVB state. (Lower panel)~Magnetization data are shown, with the same system sizes now increasing from top to bottom. The thick black lines above the {\grey} shading are the $L=\infty$ extrapolation. The magnetic order for $\mathbf{Q} = (\pi,\pi)$ and $\mathbf{Q} = (\pi,0)$ both vanish in the small region between $g_{\text{c}1} \doteq 0.54(1)$ and $g_{\text{c}2} \doteq 0.5891(3)$. }

\caption{\label{FIG:dimer_pattern} Grid lines depict the dimer correlations $C_{ijkl} = \langle (\mathbf{S}_i\cdot\mathbf{S}_j)(\mathbf{S}_k\cdot\mathbf{S}_l)\rangle$ on the nearest-{\neighbour} links $(k,l)$ of the square lattice, measured with respect to the thick, dark dimer $(i,j)$ at the {\centre}. The correlations are computed for the $L=28$ system. The {\greyscale} intensity represents correlation strength---presented as the fourth power of $(1+\tfrac{3}{2}r_{ij;kl}^{3/2})C_{ijkl}$, where $r_{ij;kl}$ is the distance measured from the {\centre} of the $(i,j)$ bond to the {\centre} of the $(k,l)$ bond. Dotted lines indicate a negative (anticorrelated) value. The top panel shows results for the NNRVB state, presented for comparison's sake. The bottom panel shows results for the energy-optimized state at $g=0.58$. In each case, a $10\times 10$ section of the full valence bond loop configuration, obtained from a snapshot of the Monte Carlo simulation, is overlaid. }

101 — 1209.6479

\caption{Maximum secure QBER as a function of $\delta_{\theta} = 90^{\circ} - \theta$. The horizontal dashed line (\textcolor{red}{S-P}) corresponds to the Shor-Preskill bound of about 0.11.}

102 — 1209.6573

\caption{\label{Fig:9Be_rot} Evidence for the ground state $K=3/2$ (\full) and excited $K=1/2$ (\dashed) rotational bands in $^9$Be from ab initio calculations with JISP16 (\opentriangle : $N_{\max}=8$, \opentriangledown : $N_{\max}=10$, \opendiamond : $N_{\max}=12$) Left: Negative parity spectrum as function of $J(J+1)$ (\fullsquare : experimental data from Ref.~\cite{Tilley2004155}.); Right: Quadrupole moments of the rotational bands normalized by $Q_0$.}

\caption{\label{Fig:Beground} Ground state energies (left) and parity splittings (right) of Be-isotopes with JISP16 (variational upperbound: \opentriangledown, NCFC: \opendiamond), together with experimental data (\fullsquare) from Refs.~\cite{Tilley2004155,Audi19971}.}

\caption{\label{Fig:14N} Ground state energies of $^{14}$N (left) and GT matrix element (right) as function of $N_{\max}$ with JISP16 at $\hbar\omega = 27.5$~MeV (\opendiamond), with chiral two-body forces at $\hbar\omega = 14.0$~MeV (\opencircle), with chiral two- and three-body forces at $\hbar\omega = 14.0$~MeV with $(c_D,c_E)=(-0.2,-0.205)$ (\fullcircle) and with $(c_D,c_E)=(-2.0,-0.501)$ (\fulldiamond), and experimental data (\fullsquare).}

\caption{\label{Fig:14Nmixed} $M_{GT}$ between the ground states of $^{14}$N and $^{14}$C, using the $^{14}$N wavefunction obtained with 3NF, but the $^{14}$C wavefunction obtained without 3NF (purple \fullcircle), and vice versa (maroon \opendiamond). For comparison, we also include the results with (red cross) and without (blue \fullsquare) 3NF for both wavefunctions.}

103 — 1210.0391

\caption{\label{fig:iso}Substeps associated with the $\sqrt{3}\times \sqrt{3}$ commensurate phase formation in the nitrogen adsorption isotherm measurements at $T=77.4$~K on our ZYX substrate~(\opencircle) with a dummy cell and those on Grafoil substrate~(\fullcircle)~\cite{niimiRSI}. The intersections of the dashed lines show the ends of the substeps. The inset additionally shows those at $T=71.8$~K~(\opensquare), $75.5$~K~$(\opentriangle)$ after mounting the ZYX substrate on our system with appropriate horizontal scaling.}

\caption{\label{fig:c}{\bf(a)} A temperature dependency of the $C_{\mathrm{add}}$. Data measured with the adiabatic heat-pulse method (red circles), those with the relaxation method (blue disks), expected addendum of nylon(\dashddot) and that of other parts(\dashed) are shown. Expected heat-capacity of 2nd-layer $^4$He(\chain)~\cite{grey}, $^3$He(\longbroken)~\cite{sv} in the 4/7 phases and 1st-layer $^4$He(\dotted) in the $\sqrt{3}\times \sqrt{3}$ phase on Grafoil and that on ZYX~\cite{root3} are also shown. {\bf(b)} A time evolution of the temperature of the addendum with the adiabatic heat-pulse method. {\bf(c)} That with the relaxation method.}

104 — 1210.2544

\caption{Two different 2-colorings of a graph with maximum degree 5 are shown in \subref{fig:introvalidexample} and \subref{fig:introinvalidexample}. The coloring in \subref{fig:introvalidexample} admits a full functional orientation, such as the one shown, on the induced subgraphs. On the other hand, the subgraphs induced by the coloring in \subref{fig:introinvalidexample} have no full functional orientation. The graph in \subref{fig:intronosolution} is not \fotwocolor able.\label{fig:example}}

105 — 1210.3053

\caption{BPT diagnostic diagram \citep*{1981PASP...93....5B}. The x-axis denotes the ratio log([\ion{N}{2}] ${\lambda6586}$/H$\alpha$) and the y-axis the ratio log([\ion{O}{3}] ${\lambda5007}$/H$\beta$). The data point marks the host galaxy of SN Primo, with the arrow denoting the upper-limit of the ratio derived from the non-detection of [\ion{N}{2}] ${\lambda6586}$. The vertical bar denotes the 1$\sigma$ error bar of the ratio. \red{For illustrative purposes we overplot the 15 emission-line hosts from the \citet{2010ApJ...722..566L} sample of SN Ia host galaxies with SDSS spectra.} \label{fig_BPT}}

\caption{ \red{The SFR-mass relation for SN Ia host galaxies. The (red) asterisk denotes the host of SN Primo with error bars. Filled (blue) circles mark star-forming hosts in each sample. The (magenta) crosses marks the passive hosts in each sample.} The solid and dotted (green) lines show the correlation between SFR and stellar mass for $z\sim0$ (bottom), $z=1$ (middle), and $z=2$ (top) from \citet[$z=2$]{2007ApJ...670..156D} and \citet[$z\sim0$ and $z=1$]{2007A&A...468...33E}. The solid section of each line marks the range of validity of the relations. (a) The low redshift sample ($z<0.21$) from SDSS \citep{2010ApJ...722..566L}. The dashed line marks the cut, $\log(sSFR)=-10.6$, between star forming and passive galaxies. The contours mark the region enclosing 68\% and 95\% of the star-forming sample. \red{(b) The high redshift samples from HST \citep[$0.95<z<1.8$, open (orange) circles]{2011ApJ...731...72T} and SNLS \citep[$0.2 < z<1.0$, filled (blue) circles / (magenta) crosses]{2010MNRAS.406..782S}. The apparent upper diagonal ridge-line for the \citet{2010MNRAS.406..782S} data is due to shortcomings in their SED fitting. %The apparent upper limit on sSFR in the \citet{2010MNRAS.406..782S} data is do to shortcomings in the SED fitting. } \label{fig_sfr_mass}}

106 — 1210.3253

\caption{(Color online) Magnetic components for the eight Co atoms: Co$_1$ (0.331, -0.081, 1/8), Co$_2$ (0.331, 0.081, 3/8), Co$_3$ (0.169, 0.081, 5/8), Co$_4$ (0.169, -0.081, 7/8), Co$_5$ (0.169, 0.419, 1/8), Co$_6$ (0.169, 0.581, 3/8), Co$_7$ (0.331, 0.581, 5/8), and Co$_8$ (0.331, 0.419, 7/8), for each irreducible representation $\tau^1$, $\tau^2$, $\tau^3$, and $\tau^4$. The non-circled signs ('\textcolor{red}{+}' and '\textcolor{red}{$-$}') and the circled signs ('\textcolor{blue}{$\oplus$}' and '\textcolor{blue}{$\ominus$}') correspond to independent magnetic moments.}

\caption{(Color online) (a) Temperature dependence of the magnetic signal: neutron counts on top of the strong 2\,$\bar{3}$\,0 reflection of the N\'eel phase for various magnetic field values between 0 and 3.25~T and on top of the 2\,$\bar{3}$\,0.127 reflection of the LSDW phase at$H=5$ T. The latter curve was obtained from $Q_L-$scans performed across the 2\,$\bar{3}$\,0.127 position counting one minute per point (and then renormalized to 10 seconds for the figure). The black solid lines are power-law fits performed on a temperature range$\delta T=T_c-T \sim 1$~K as explained in the text. (b) and (c) The set of the $\beta$ and $T_c$ values as obtained for various fitting temperature ranges $\delta T$ at $H=0$ together with a linear extrapolation of both parameters to $\delta T=0$ for a precise determination of their values (see text). Inset of panel (a): log-log plot of the magnetic intensity with subtracted background, $I-I_0$, as a function of the reduced temperature, $1-T/T_c$, at $H=0$. The data (open circles) and the power-law fit (solid line) are presented using the extrapolated values for $\delta T \rightarrow 0$: $T_c = 5.556$~K (for the data and the fit) and $\beta=0.307$ (for the fit).}

\caption{(Color online) Field dependence of the magnetic signal at various temperatures between 1.85~K and 5~K: neutron counts on top of the strong $\bar{2}$\,3\,0 reflection of the N\'eel phase. The black solid lines are fit to a power law as explained in the text. Inset:$\beta'$ values as a function of the temperature (see text). The dashed black line is a linear fit of these data.}

\caption{(Color online) $Q_L-$scans performed across the ${\bf Q}=(\bar{2},\bar{3},0)$ scattering vector at $T=50$~mK for different increasing magnetic field values. (a) N\'eel phase, (b) Coexistence of the N\'eel and LSDW phases, (c) and (d) LSDW phase. For a better visibility, the scans measured at 4 and 3.98~T in (b) were shifted vertically by 2000 and 4000, respectively.}

\caption{(Color online) Magnetic $H$-$T$ phase diagram of BaCo$_2$V$_2$O$_8$ obtained from the specific heat measurements presented in Sec.~\ref{subsec:IIIA} (black squares) and from the present neutron diffraction measurements (red circles). As concerns the N\'eel-LSDW phase transition (see enlargement in the inset), the solid and open red circles correspond to increasing and decreasing field scans, respectively. The blue open triangles in the main panel are the data of Kimura{\it et al.}~\cite{Kimura2008} obtained from macroscopic measurements and the blue dashed line is the theoretical LSDW-paramagnetic transition line predicted by Okunishi and Suzuki.\cite{Okunishi2007}}

\caption{(color online) Magnetic structure of \bacovo in the LSDW phase at $H=4.2$ T (decreased from 6~T) for the magnetic domain~\#1, shown over about one half period ($\sim 5 c$), using the same labeling and the same colors as in Fig.~\ref{StrucMagH0} for the two types of cobalt chains. This projection on the $(b,c)$ plane is to be compared to Fig.~\ref{StrucMagH0}(a), sketching the magnetic structure of domain \#1 in zero field. The magnetic domain~\#2 of the LSDW phase can simply be obtained from domain\#1 by reverting all spins plotted in red (like for the structure in zero magnetic field).}

107 — 1210.3600

\caption{ \label{fig:map2} The four columns show the logarithm of Ly$\alpha$ surface brightness maps (in units of $\SBunit$) for four randomly selected large galaxies of virial masses both exceeding $10^{12}\msun$ at $z=3.1$ with the primary galaxy centered on their respective panel. For each column the bottom panel is obtained, if one only includes galaxies within $\pm R_{\rm vir}$ of the primary galaxy along the line of sight, where $R_{\rm vir}$ is the virial radius of the host halo of the primary galaxy. The bottom panel is obtained, including all galaxies within $\pm 10h^{-1}$Mpc (comoving) of the primary galaxy along the line of sight. The length shown is in physical kpc. The galaxies are taken from the simulation with resolution of $110h^{-1}$pc \citet[][]{2011eCen}. The effects of dust and faint sources have not been included yet in these plots (see the text for more details). {\color{blue} \bf ZZ: should make the x- and y-labels, and axis tick labels larger -- they are hard to see.} }

108 — 1210.3948

\caption{Modal density of the Atlas piano soundboard. Dots: 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{$\ast$}}), in Fig.\ref{fig:pianoatlas}. The estimated values are the reciprocal of the moving average of six successive modal spacings, reported at the mid-frequency of the whole interval.}

109 — 1210.4671

\caption{ \label{fig_3} (Color online) Estimates for the prefactor of the corner contribution $\left\langle {\cal S}_{\square}^{cr}\right\rangle$ of the square subsystem obtained from two-point fits. The asymptotic values as $L\to\infty$ are the same for site and bond percolation and agree with the conformal prediction. Inset: Large scale topology of the clusters. The square is divided into four quadrants and the different topologies of clusters in the quadrants are illustrated: empty box $\to$ no site; filled box $\to$ at least one occupied site.} \end{figure} %%%%%%%%%% FIG 3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Making use of the fact that asymptotically $\left\langle{\cal S}_{\square}^{cr}(L)\right\rangle=b \ln L + \text{const}$, we have calculated effective, size-dependent estimates for the prefactor by two-point fits, by comparing the average corner contributions in finite systems of size $L$ and $2L$. The results are presented in Fig.\ref{fig_3} both for site and bond percolations. With increasing $L$ the effective prefactors approach a common, universal limiting value of $b=-0.077(1)$. This is to be compared with the conformal result in Eqs. (\ref{N_Gamma},\ref{cardy_peschel}) and (\ref{c'}) : $-C'(1)=-5 \sqrt{3}/(36 \pi)=-0.07657$, thus the agreement is satisfactory. We note that a previous numerical estimate in Ref.~\onlinecite{yu07} for smaller systems has obtained: $b=-0.06(1)$. The corner contribution is related to the large scale topology of the clusters, as illustrated in the inset of Fig.\ref{fig_3} and discussed in Sec.\ref{sec:corner_pr}. We have also studied the $p$-dependence of the corner contribution to $\left\langle{\cal S}_{\square} \right\rangle$ outside the critical point by the geometrical approach. As can be seen in Fig.\ref{fig_4} ${\cal S}_{\square}^{cr}$ has a peak around $p=p_c$ and close to $p_c$ the extrapolated curve can be well described by the scaling result: % \be{\cal S}_{\square}^{cr}(p) \simeq b' \ln(p_c-p)+const\;,\label{S_corner_p} \ee % where $b'=b \nu$, with $\nu=4/3$ being the correlation length critical exponent for percolation. Indeed, assuming the form in Eq.(\ref{S_corner_p}) we have calculated effective, $p$-dependent prefactors by two-point fits, which are shown in the inset of Fig.\ref{fig_4}. The extrapolated value for $p \to p_c$ ($0.106(10)$) is consistent with the scaling prediction ($0.102$). %%%%%%%%%% FIG 4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_4} \end{center} %\vskip -.5cm \caption{ \label{fig_4} (Color online) $\Delta p=p-p_c$-dependence of the corner contribution to ${\cal S}_{\square}$ for different sizes, $L=16,32,\dots,2048$ from below for bond percolation. Close to $p_c$ the extrapolated curve has a logarithmic singularity, see Eq.(\ref{S_corner_p}). In the inset the effective prefactors, $b'$, are shown as a function of $|\Delta p|$. } \end{figure} %%%%%%%%%% FIG 4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Sheared square subsystem} \label{sec:sheared_square} %%%%%%%%%% FIG 5 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_5} \end{center} %\vskip -.5cm \caption{ \label{fig_5} (Color online) Estimates for the prefactor $-b$ for sheared squares as a function of $\ln(\gamma)$. The numerical results are for $L=128,~256,~\dots,~4096$ from below, the full line represents the conformal result in Eq.(\ref{sheared}). In the inset the ratio of the extrapolated numerical data and the conformal result is given.} \end{figure} %%%%%%%%%% FIG 5 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% The square subsystem used in the previous subsection is sheared now to a parallelogram, having the same surface: $\ell^2=L^2/4$ and its (smaller) angle is $\gamma$. The geometrical approach to calculate the corner contribution to ${\cal S}_{\Diamond}^{cr}$ can be extended in this case for specific values of $\gamma$ given by the condition: $\tan(\gamma)=1/n$, with $n=0,1,\dots,\infty$ being an integer. As shown in Fig.\ref{fig_5} with increasing $L$ up to $L=4096$ the numerical results approach the conformal prediction: % \be C'(1)=-\frac{c'(1)}{12}\left[4-\pi\left(\frac{1}{\gamma}+\frac{1}{\pi-\gamma}+\frac{1}{\pi+\gamma}+\frac{1}{2\pi-\gamma}\right)\right]\;. \label{sheared} \ee % Performing an extrapolation with an $\ln L/L$ correction term we have obtained an excellent agreement, as shown in the inset of Fig.\ref{fig_5}. For other values of $\gamma$, which do not fit to the geometrical approach we have made calculations by the difference method. Also in these cases the numerical results are found to agree with the conformal prediction. \subsection{Anisotropic percolation} \label{sec:anisotrop} Here we consider anisotropic bond percolation, in which the probabilities are $p_x$ (horizontally) and $p_y$ (vertically) and the critical point is given by the condition: $p_x+p_y=1$. The system is a diagonally placed square with $2L^2$ sites, and the subsystem is also a diagonally oriented square, having $L^2/2$ sites and its boundary contains $4 \times L/2$ sites, see the third figure in the first row of Fig.\ref{fig_1}. This anisotropic system with symmetric angles, $\pi/2$, is conjectured to be equivalent in the scaling limit to an isotropic system with asymmetric angles, $\gamma$ and $\pi-\gamma$, such that % \be \dfrac{p_y}{p_x}=\dfrac{\sin((\pi-\gamma)/3)}{\sin(\gamma/3)}\;.\label{p_y/p_x} \ee % This follows from the requirement that there exists a discretely holomorphic observable\cite{ikhlef_cardy}. More recently Grimmett and Manolescu\cite{grimmett} have proved that many properties of the scaling limit are the same if a more general inhomogenously anisotropic lattice is embedded in the plane according to this prescription. %%%%%%%%%% FIG 6 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_6} \end{center} %\vskip -.5cm \caption{ \label{fig_6} (Color online) Estimates for the prefactor $-b$ for anisotropic percolation for finite systems with $L=32,~64,~\dots,~1024$ from below. The opening angle of the equivalent sheared square, $\gamma$ is given by Eq.(\ref{p_y/p_x}) and the full line represents the conformal result. In the inset the ratio of the finite-size results and the conformal results are given.} \end{figure} %%%%%%%%%% FIG 6 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In the anisotropic system we have chosen the probabilities in such a way, that the corresponding angle, $\gamma$, has satisfied the condition: $\tan(\gamma)=1/n$, with $n=0,1,\dots,\infty$. In this way we could directly compare the results for anisotropic systems with those obtained for sheared squares in Sec. \ref{sec:sheared_square}. The numerical results for the prefactor obtained on finite systems up to $L=1024$ are are shown in Fig.\ref{fig_6}. For the three largest systems there is no systematic size-dependence of our data, the error is purely statistical and the numerical results agree well with the conformal prediction. \subsection{Corner probability} \label{sec:corner_pr} For the (sheared) square subsystem we used the geometrical method, in which the $L \times L$ system is divided for four $L/2 \times L/2$ squares and for four $L/2 \times L$ stripes and the difference in the number of crossing clusters in Eq.(\ref{S_corner}) is just the corner contribution. Having a general (connected) cluster it could occupy different parts of the four quadrants and its topology could be of four different types as illustrated in the inset of Fig.\ref{fig_3}. Among these the I, II and IV type of topology gives identical contribution both for stripes and squares. Clusters with topology III, however, are crossing clusters for all the four possible stripes, but these are crossing clusters only for three out of four squares. (In the left-bottom square there is no crossing.) Let us denote the occurrence probability of type III cluster as $P_3(L)$, which is given by the ratio of such clusters in a $L \times L$ square, which have points in three quadrants, but have no point in the fourth one. As argued above $P_3(L)$ is proportional to the corner contribution of crossing squares, more precisely: % \be P_3(L)=\dfrac{1}{4}{\cal S}_{\square}^{cr}(L)=-\dfrac{1}{4}C'(1)\ln L +\text{const.}\;.\label{P_3} \ee % This results is valid for sheared squares with angle $\gamma$, as well as for anisotropic percolation with a square subsystem. In these cases the appropriate results of ${\cal S}_{\square}^{cr}(L)$ have to be used. Interestingly this corner probability has a logarithmic $L$-dependence and its prefactor is known exactly. \subsection{Other subsystem geometries} \label{sec:Other} We have also studied subsystems with different geometries: equilateral triangle, circle and section of a circle, these are illustrated in the second row of Fig.\ref{fig_1}. In these cases the largest linear scale of the subsystem is fixed to $L/2$, while an angle was varied. In the calculations the difference approach in Eq.(\ref{delta_S}) has been used. \subsubsection{Equilateral triangle} %%%%%%%%%% FIG 7 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_7} \end{center} %\vskip -.5cm \caption{ \label{fig_7} (Color online) Prefactor of the logarithm for an equilateral triangle as a function of the base angle $\gamma$ compared with the conformal result in Eq.(\ref{triangle}). The numerical results are extrapolated up to $L=4096$ for site percolation. The statistical error is larger, than the difference between the conformal and the numerical results.} \end{figure} %%%%%%%%%% FIG 7 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% For an equilateral triangle with a base angle $\gamma$ the extrapolated results for $-b$ are shown in Fig.\ref{fig_7} , which is compared with the conformal prediction: % \be C'(1)=-\frac{c'(1)}{24}\left[6-\pi\left(\frac{2} {\gamma}+\frac{1}{\pi-2\gamma}+\frac{1}{\pi+2\gamma}+\frac{2}{2\pi-\gamma}\right)\right]\;. \label{triangle} \ee % As seen in Fig. \ref{fig_7} there is a satisfactory agreement, although the statistical error of the numerical results is comparatively large, in particular for small and large angles. \subsubsection{Circle and section of a circle} %%%%%%%%%% FIG 8 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_8} \end{center} %\vskip -.5cm \caption{ \label{fig_8} (Color online) Prefactor of the logarithm $-b$ for a section of a circle as a function of the angle $\gamma$. The full line represents the conformal result in Eq.(\ref{section}). The differences between the conformal and numerical results are of the order of the statistical error. In the inset the corner contribution is shown for a circle: there is no systematic radius dependence. The numerical results are extrapolated up to $L=4096$ (in the main figure) and up to $L=8192$ (in the inset) for site percolation.} \end{figure} %%%%%%%%%% FIG 8 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% According to the conformal results, there is no logarithmic correction to $S_{\bigcirc}$ for a circle shaped subsystem. Indeed the numerical results in the inset of Fig.\ref{fig_8} are in agreement with this statement: $\Delta S_{\bigcirc}$ approaches a finite limiting value of $0.65(5)$. In contrast, for a section of a circle with an angle $0 < \gamma < \pi$ there is a logarithmic correction, the prefactor of which is given by conformal invariance: % \be C'(1)=-\frac{c'(1)}{12}\left[2-\pi\left(\frac{1}{\gamma}+\frac{1}{2\pi-\gamma}\right)\right]\; \label{section} \ee % The numerical results in Fig.\ref{fig_8} are in agreement with the conformal prediction, although the statistical error of the numerical results is comparatively large. \subsection{Line segment} \label{sec:line} According to the derivation in Sec.\ref{sec:Potts} $\Gamma$ does not have to be a closed curve. For example it could be a straight line of length $\ell$. In that case we have only a corner contribution from two exterior angles, each $\gamma=2 \pi$, so that $C_{\Gamma}(Q)=-c(Q)/8$. To study this problem we can use the geometrical approach, when the line is oriented parallel with one of the axes of the square lattice. Then the corner contribution is related to the difference between the cluster numbers obtained for a periodic line of length $L$ and that of two segments of lengths $\ell=L/2$. In this case the corner contribution is simply half of the number of common clusters between the two line segments. %%%%%%%%%% FIG 9 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[!ht] \begin{center} \includegraphics[width=3.2in,angle=0]{Fig_9} \end{center} %\vskip -.5cm \caption{ \label{fig_9} (Color online) $\Delta S$ for the line segment obtained by the difference approach for different lengths, $L$. In the inset the $p$-dependence of $S^{cr}$ is shown close to the critical point for finite systems with $L=64, 128, 256$ and $512$ from below.} \end{figure} %%%%%%%%%% FIG 9 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In the numerical calculation we have used the difference approach for a line segment of length $\ell=L/2$, which is placed in random positions and random orientations with respect to the axes of the periodic square lattice of size $L \times L$. The average of $\Delta S_{\arrowvert}(L)$ is shown in Fig. \ref{fig_9} as a function of $\ln L$. The logarithmic dependence is clearly visible with a prefactor estimated as $b=0.086(1)$, which agrees fairly well with the conformal prediction: $-C'_{\Gamma}(1)=\frac{5\sqrt{3}}{32\pi}\approx 0.08615$. We have also checked the $p$-dependence of $S^{cr}_{\arrowvert}(p)$, which is shown in the inset of Fig. \ref{fig_9}. A singularity is developed at the critical point as $L \to \infty$. We note that if the line segment is put to the open boundary of the system, then the prefactor is different, being $b=\frac{\sqrt{3}}{4\pi}\approx 0.1378$ as obtained by conformal invariance\cite{cardy01,yu07} and numerically $b=0.15(2)$ \cite{yu07}. \section{Discussion} \label{sec:disc} In this paper we considered the logarithmic terms in the mean number $N_\Gamma$ of clusters in critical percolation which intersect a curve $\Gamma$ in the cases when it has sharp corners or end points. We have shown that the Cardy-Peschel result\cite{cardypeschel} can be simply applied to compute the universal coefficients of these logarithmic terms, and that accurate numerical estimates agree very well with this, for a variety of shapes for $\Gamma$. We also considered anisotropic bond percolation on the square lattice and showed that if Eq.~(\ref{p_y/p_x}) is used to deform the lattice then the predictions again agree with numerics if the correct effective corner angle is used. We have also pointed out a relation between the corner contribution to $N_{\Gamma}$ and the statistics of cluster shapes. Our study is related to recent investigations of shape dependent terms of different thermodynamic quantities, mainly the free energy, of $2d$ critical systems. For critical percolation, however, there is no logarithmic corner contribution to the free energy, since the central charge is $c(Q=1)=0$. In the latter problem the corner contribution to the cluster numbers and its higher moments are of interest, which are related to derivatives of $c(Q)$ at $Q=1$. The excellent agreement of the numerical data with the theory serves as confirmation that critical percolation is indeed conformally invariant, of the Coulomb gas predictions for $c'(Q=1)$, and of the formula (\ref{p_y/p_x}). Extensions are obviously possible: for example by taking further derivatives with respect to $Q$ we can find results for the corner contributions to the higher moments of the distribution of the cluster numbers $N_\Gamma$. We note that recently\cite{Vasseur} an explicit example has been found of a correlation function in percolation which contains a multiplicative logarithm. As with our result, this may be understood\cite{cardylogs} from the necessity of having to take a suitable derivative with respect to $Q$ at $Q=1$ in the Potts model. \begin{acknowledgments} This work has been supported by the Hungarian National Research Fund under Grants No. OTKA K75324 and K77629. This work has been partly done when two of the authors (J. C. and F. I.) were guests of the Galileo Galilei Institute in Florence whose hospitality is kindly acknowledged. \end{acknowledgments} \begin{thebibliography}{99} \vskip -.5cm \bibitem{smirnov} S. Smirnov, C. R. Acad. Sci. Paris S\'er. I Math., 333 , no. 3, 239-244 (2001); J. Tsai, S.C.P. Yam, W. Zhou, arXiv:1112.2017.\bibitem{sle} O. Schramm, Israel J. Math. 118, 221 (2000); S. Smirnov and W. Werner, Math. Research Letters 8, no. 5-6, 729-744 (2001). \bibitem{Fortuin-Kasteleyn} P. W. Kasteleyn and C. M. Fortuin, J. Phys. Soc. Japan \textbf{26}, 11 (1969). \bibitem{cardypeschel} J. Cardy and I. Peschel, Nucl. Phys. B, 300 [FS22], 377 (1988). \bibitem{note} Corner contributions to finite-size corrections of critical percolation involve only the interior parts of Eq.(\ref{cardy_peschel}). For a rectangle it has been calculated by P. Kleban and R.M. Ziff, Phys. Rev. B \textbf{57}, R8075 (1998). \bibitem{CG} See for example J. Cardy, Ann. Phys., 318(1), 81 (2005). \bibitem{lin07} Y-C.Lin, F. Igl\'oi and H. Rieger,\prl \textbf{99}, 147202 (2007). \bibitem{yu07} R. Yu, H. Saleur and S. Haas, \prb \textbf{77}, 140402 (2008). \bibitem{senthil_sachdev} T. Senthil and S. Sachdev, Phys. Rev. Lett. \textbf{77}, 5292 (1996). \bibitem{sedgewick} We applied the weighted union-find algorithm with path compression, see for example R. Sedgewick, {\sl Algorithms}, 2nd edition, Addison-Wesley, Reading, Mass. (1988). \bibitem{kovacs_igloi12} I. A. Kov\'acs and F. Igl\'oi, EPL\textbf{97}, 67009 (2012). \bibitem{ikhlef_cardy} Y. Ikhlef and J. Cardy, J. Phys. A: Math. Theor. \textbf{42}, 102001 (2009). \bibitem{grimmett} G. R. Grimmett, I. Manolescu, arXiv:1108.2784. \bibitem{cardy01} J. Cardy, Phys. Rev. Lett., 84, 3507 (2000). \bibitem{Vasseur} R. Vasseur, J. L. Jacobsen and H. Saleur, J. Stat. Mech. L07001 (2012). \bibitem{cardylogs} J. Cardy, arXiv:cond-mat/9911024; J. Cardy, in {\sl Statistical Field Theories}, A. Cappelli and G.Mussardo eds., Kluwer (2002). \end{thebibliography} \end{document} }

110 — 1210.4769

\caption{Broadband SED of Tycho's SNR with the observed data in radio \citep{Kothes2006}, X-rays \citep{Tamagawa2009} and \grays\(\Fermi : \citealt{Giordano2012}; \veritas : \citealt{Acciari2011}).}

111 — 1210.4800

\caption{ A list of frequently studied $\beta\beta$ isotopes and their Q-values. $\gamma$-production cross sections or upper limits for important transitions in $^{nat}$Cu are given for neutrons between 2.897 and 4.196 MeV. Where the cross section is listed as NA, the experimenters were unable to place a limit due the \gray being outside the range of the detection system.\label{ana:tab:fep} }{ \begin{tabular}{@{}cccc@{}} \hline $\beta\beta$ isotopes & $\gamma$ ray & SEP & DEP\\ & [mb] & [mb] & [mb] \\ \colrule $^{48}$Ca & \fcu NA & \fcu NA & \fcu NA \\ $^{76}$Ge & \fcu $<$0.388(3) NA & & \\ $^{82}$Se & \tcu 9.42(32) & $<$0.324(3) & \tcu NA\\ $^{96}$Zr & \fcu 1.12(2) & $<$0.241(3) & \fcu NA\\ $^{100}$Mo &\tcu 9.42(32) & \tcu 0.59(22) & \tcu NA\\ $^{116}$Cd & \tcu 4.41(23) & & \\ $^{130}$Te & \tcu 9.42(32) & \tcu 9.42(32) & \tcu $<$0.316(3) \\ $^{136}$Xe & \tcu 0.62(10) & $<$ 0.392(3) & \tcu 1.03(10) \\ $^{150}$Nd & $<$ 0.319(3) & $<$ 0.265(3) & \fcu NA\\ \hline \end{tabular} }

\caption{The $\gamma$ production cross sections for 2977, 3032, and 3042-keV \grays in \tcu. These \gray are all ground state transitions, and therefore the energy of the \gray corresponds to the energy level in \tcu. These \gray transitions are in the vicinity of the $^{82}$Se and $^{100}$Mo endpoint energies. The corresponding levels for these \grays are not included in the ENDF/B-VII evaluation for \tcu.}

\caption{The $\gamma$ production cross section for 3355-keV \gray in \fcu. This $\gamma$ ray transition is in the vicinity of the $^{96}$Zr endpoint energy, and is a ground state transistion from the 3355 energy level in \fcu. The corresponding level for this \gray is not included in the ENDF/B-VII evaluation for \fcu.}

\caption{The $\gamma$ production cross section for 2808-keV \gray in \tcu. This $\gamma$ ray transition is in the vicinity of the $^{116}$Cd endpoint energy. The corresponding level excitation cross section for this transition is shown in Fig. \ref{fig:ana:cdfep} and is included in the ENDF/B-VII evaluation for \tcu.}

\caption{The $\gamma$ production cross section for 2535-keV \gray in \tcu. This $\gamma$ ray transition is in the vicinity of the $^{130}$Te endpoint energy. The corresponding level excitation cross section for this transition is shown in Fig. \ref{fig:ana:tefep} and is included in the ENDF/B-VII evaluation for \tcu.}

\caption{The $\gamma$ production cross section for 3429-keV \gray in \tcu. This $\gamma$ ray transition is in the vicinity of the $^{136}$Xe endpoint energy, and is a ground state transition from the 3429-keV energy level in \tcu. The corresponding level for this \gray is not included in the ENDF/B-VII evaluation for \tcu.}

\caption{The $\gamma$ production cross section for 3476-keV \gray in \tcu. This $\gamma$ ray transition is in the vicinity of the $^{136}$Xe double-escape peak energy region, and is a ground state transition from the 3476-keV energy level in \tcu. The corresponding level for this \gray is not included in the ENDF/B-VII evaluation for \tcu}

112 — 1210.5199

\caption{(Color online) (Left) Invariant yields vs. \pt \of non-photonic electron at\saaTwoHundred, and scaled STAR published $p+p$ \cite{starpp200}. Error bars and boxes are statistical and systematic errors, respectively. FONLL predictions are scaled by $N_{coll}$ shown as curves. (Upper-right) Non-photonic electrons nuclear modification factor, $R_{AA}$, at \saaTwoHundred\compared to models\cite{dglv}-\cite{ adscft}. Grey band is the light hadrons $R_{AA}$. Error bars and brackets are Au+Au statistical and systematic errors, respectively. Error boxes are the uncertainties from our baseline $p+p$ measurement. (Lower-right) Non-photonic electrons azimuthal anistropy $v_2\{2\}$, $v_2\{4\}$ and $v\{EP\}$ at \saaTwoHundred. Error bars and brackets are statistical and systematic errors, respectively.}

113 — 1210.7268

\caption[]{Rotation period evolution from the main sequence to the giant branch as a function of the (J -- H)$_{C3}$ index. The models are shown for [Fe/H] = 0 and 1 M$_{\odot}$. The dashed line represents the rotation period evolution for a solid-body model. The solid line represents the computed model with differential rotation. Open circles represent subgiants with masses of about one solar mass and rotational periods determined from the rotation-activity relation by \blue{\citet{Lovis_2011}}. Asterisk as in Fig~\ref{Prot_JH}.}

114 — 1210.8341

\caption{ Evolution of the structure of (LaS)$_{1.196}$VS$_2$ between 300K and 90K (a) Temperature dependence of LaS and VS$_2$ layers in-plane cell parameters. (b) Temperature dependence of the shortest and the longest vanadium-vanadium distance $d_{minV-V}$ and $d_{maxV-V}$. (c) \textcolor{red}{Comparison of V-V distances observed within (plain lines) and between (dashed lines) some representatitve vanadium tetramers formed at 300 and 90K.} }

115 — 1210.8442

\caption{Sample trajectory of $\theta_i^{(t)}$ in \textbf{VarO} and \textbf{SemiO}. In all above, \textcolor[rgb]{0.00,0.00,1.00}{Blue} line is for \textbf{VarO}, \textcolor[rgb]{0.00,0.50,0.00}{Green} line is for \textbf{SemiO}, and \textcolor[rgb]{1.00,0.00,0.00}{Red} line is the moving average of Green line over window size 30.}

\caption{Sample trajectory from \textbf{SemiU}. Each plot shows four neurons resulted from two splits on a single neuron. Thick dash lines: neurons with positive outgoing weights. Thin solid lines: corresponding neurons with negative outgoing weights. \textcolor[rgb]{0.00,0.00,1.00}{Blue} and \textcolor[rgb]{1.00,0.00,0.00}{Red} lines are for the two neurons from splitting one sampling into two. \textcolor[rgb]{0.00,0.50,0.00}{Green} line is the average of Blue and Red. Ideally, lines with same color should be equal, and green lines should be at 0.5.}

\caption{Some statistics of the algorithms. (a) means of trajectories in \textbf{SemiO} vs. converged activation values in \textbf{VarO} for the same hidden variables. (b) standard deviation vs. mean, from the trajectories in \textbf{SemiO}. (c) histogram of standard\_deviation(red line + blue line - 1.0). (d) histogram of standard\_deviation(dash line - solid line). (Here red, blue, dash, solid lines refer to figure~\ref{fig:semiu}.)}

116 — 1211.0370

\caption{(Color online). Experimental test of complementarity relations. Subfigures (a) and (b) \blk correspond to the simple and the optimal estimates of $X$ described in the text. Each subfigure \blk shows, in descending order, the left hand sides of the \red Ozawa and Hall relations (\ref{ozawa}) and (\ref{hall}) \blk[indistinguishable in (a)]; the new complementarity relation (\ref{hall2}); and the left hand side of the Arthurs-Kelly relation (\ref{ak}). The pink shading indicates the region corresponding to violation of any of these relations. Error bars not shown are smaller than the size of the markers. Solid curves are theoretical predictions. }

117 — 1211.0533

\caption{ Periodic table of topological insulators and superconductors for the 10 symmetry classes in various spatial dimensions. Topological insulators (superconductors) in the complex symmetry classes (A and AIII) are related to the chiral U(1) anomaly. The primary series of the topological insulators (superconductors) with an integer ($\mathbb{Z}$) classification in the eight real symmetry classes are located on the diagonal in the table. In even space-time dimensions (odd space dimensions) they are predicted from the chiral anomaly in the presence of background gravity (\textcolor{blue}{$\mathbb{Z}^{\diamondsuit}$}), and from the chiral anomaly in the presence of both background gravity and U(1) gauge field (\textcolor{green}{$\mathbb{Z}^{\clubsuit}$}). The topological response of topological phases in odd space-time dimensions (\textcolor{red}{$\mathbb{Z}^{\heartsuit}$} and \textcolor{magenta}{$\mathbb{Z}^{\spadesuit}$}) follows from their higher-dimensional ancestor (\textcolor{blue}{$\mathbb{Z}^{\diamondsuit}$} and \textcolor{green}{$\mathbb{Z}^{\clubsuit}$}), respectively. }

118 — 1211.1323

\caption{Learning curves of the real data set: sensitivities for recognition of red blood cells (rbc), leukocytes (leu), and BT-20 breast tumour cell line (bt). % Black: Sensitivity observed for 100 iterations of 5-fold cross validation on the complete data set, approximating the best possible performance of a 10 latent variable PLS-LDA on this data set. % Lines give the average, the shaded area covers the 5\tes to 95\tes percentile of iterations (bottom and middle row) and small data sets (top row). Thin lines: average ``one set large'' and ``one set cv'' (cross validation) performance are repeated in the rows above for easier comparison. % Colours: {\color{darkblue}blue} performance measured with large test set, {\color{red}red} performance measured by iterated cross validation. % Bottom row: Learning curve of \emph{one} growing data set, measured with 100$\times$ iterated 5-fold cross validation. % Middle row: The same models as in the bottom row, but performance measured with large test set. The percentiles depict the instability of the surrogate models trained during iterated cross validation, but are subject only to low uncertainty due to the finite \emph{test} sample size. % Top row: sensitivity achieved for 100 different small data sets of size \ntrain, measured with the large test set. }

119 — 1211.1779

\caption{\textcolor{black}{(Color online) Entanglement} $\Delta_{g,ent}$ plotted versus the squeezing parameter $r$ for $n_{0}=0,\ 5,\ 10,\ 50$, where $n_{0}$ is initial occupation number of the oscillator. Entanglement between the oscillator and pulse is observed when $\Delta_{ent}<1$. Strong entanglement occurs when $\Delta_{ent}\rightarrow0$. The optimal $g$ to minimize $\Delta_{g,ent}$ is shown in the inset. The results indicate presence of entanglement, even for large $n_{0}$. \label{fig:asymmetric_ent_sum_product} \textcolor{red}{ }}

\caption{(Color online) Detecting an EPR paradox between two mechanical oscillators prepared using the scheme of Eqs. (\ref{eq:m2}) with $r=r'$: (a) The threshold squeeze parameter $r_{epr}$ for observation of the EPR paradox $E_{m2|m1}<1$, versus the thermal occupation number of the oscillator $m2$. The curves are for $n_{m1}=0$ (solid), $n_{m1}=1$ (dashed), $n_{m1}=1.5\times10^{6}$ (dotted) \textcolor{red}{}(b) The threshold squeeze parameter $r_{epr}$ for observation of the EPR paradox $E_{m1|m2}<1$, versus the thermal occupation number of the oscillator $m2$. The curves are for $n_{m1}=0$ (solid), $n_{m1}=1$ (dashed), $n_{m1}=1.5\times10^{6}$(dotted). \textcolor{red}{}\textcolor{black}{{} (c)} The threshold squeeze parameter $r_{epr}$ for observation of the EPR paradox $E_{m1|m2}<1$, versus the thermal occupation number of the oscillator $m1$. The curves are for $n_{m2}=0$ (solid), $n_{m2}=10$ (dashed), $n_{m2}=100$ (dotted)\label{fig:rth}}

120 — 1211.2093

\caption{(Color online) Number of atoms $a(k)$ as measured in absorption imaging at the end of the evaporation sequence versus the dispersive interrogation pulse $d_{10}(k)$ ({\large$\bullet$}); red line is a least squares fit to the data, which is used to define $d_{\rm target}$ from a target atom number. The dataset corresponding to Fig.~\ref{fig:Hist2} ({\scriptsize\textcolor{raw}{$\blacksquare$}}) is decimated ({\scriptsize\textcolor{filtered}{$\blacksquare$}}) by excluding points outside a 10\% band around $d_{\rm target}$.\label{fig:corrplot}}

121 — 1211.2228

\caption{\textbf{Cavity spectroscopy and displacement calibration.} \textbf{a}, Cavity spectroscopy for varying spectroscopy power. The cavity excitation is measured by mapping the population left in the ground state to the qubit using a photon number state selective $\pi$ pulse. For the lowest power (\textcolor{BlueGreen}{$\bullet$}) the cavity is only excited on the $\ket{0} \rightarrow \ket{1}$ transition. As the power is increased ($\textcolor{BlueGreen}{\bullet}\rightarrow\textcolor{Dandelion}{\bullet} \rightarrow \textcolor{ForestGreen} {\bullet} \rightarrow \textcolor{CadetBlue}{\bullet}\rightarrow \textcolor{Red}{\bullet}$) different transitions appear in the spectroscopy. These transitions are n-photon transition from $\ket{0}$ to $\ket{n}$ with $n=1, 2, 3$ labeled from the right peak to left. The separation between the peaks is given by $0.5\, K/2\pi= 163$~kHz. The dashed lines are fit to the data using a multipeak Lorentzian function. \textbf{b}, Photon number probability of the Fock states $n=0\dots7$ ($\textcolor{Red}{\bullet}, \textcolor{RedOrange} {\bullet}, \textcolor{BurntOrange}{\bullet}, \textcolor{LimeGreen}{\bullet},\textcolor{JungleGreen}{\bullet},\textcolor{RoyalBlue}{\bullet},\textcolor{RubineRed}{\bullet},\textcolor{BlueViolet}{\bullet}$) for a 10~ns displacement with amplitude $|\alpha|$ of the cavity ground state. The axes were scaled by fitting a Poisson distribution with two free fit parameters to all seven Fock state populations simultaneously. The dashed lines are given by a Poisson distribution $P(|\alpha|) = |\alpha|^{2n} e^{-|\alpha|^2}/n!$ for the Fock state n.}

122 — 1211.3853

\caption{\leg{(a):} Distributions of short samples of wind component speed $u$, rescaled by their standard deviation $\sigma$ computed over 4~hr time periods. Black line indicates a Gaussian function of standard deviation $1$. \leg{(b):} Distribution of each sample standard deviation $\beta=\sigma^{-2}$ and its fit by a Gamma distribution. \leg{(c):} Superstatistical distribution corresponding to Eq.~(\ref{eq:superstatistics_4}). Color codes for varying parameters $m$ and $n$. \leg{(d):} Wind speed component distributions from the SIRTA observatory measurements (\textcolor{red}{$\bigcirc$}, $u$, and \textcolor{blue}{$\times$}, $v$) and their fit by the superstatistical distribution corresponding to Eq.~(\ref{eq:superstatistics_4}) with parameters, $m=9$, $n=0.019$ (black dashed line).}

\caption{\leg{(a):} Superstatistical distribution corresponding to Eq.~(\ref{eq:superstat_5}). Color codes for varying $m$ and $n$ parameters. \leg{(b):} Wind speed distribution from the SIRTA observatory measurements (\textcolor{red}{$\bigcirc$}, $u$, and \textcolor{blue}{$\times$}, $v$) and its fit by the superstatistical distribution given by Eq.~(\ref{eq:superstat_5}) with parameters $m=9$, $n=0.019$ (magenta dashed line).}

123 — 1211.3901

\caption{Examples of error modes in the Kinect tracker. Right (\textcolor{yellow}{$\bullet$}) and left (\textcolor{green}{$\bullet$}) hand, and shoulder (\textcolor{red}{$\bullet$}) pose estimates are marked.}

\caption{Trajectories for all 4 signers for signs 'Smaka' ('Taste') and '5' : Signer A ($\bullet$), Signer B (\textcolor{blue}{$\bullet$}), Signer C (\textcolor{yellow}{$\bullet$}), Signer D (\textcolor{green}{$\bullet$})}

124 — 1211.4425

\caption{(Color online) (a) $\mathcal M_1$ monomers at position $\vec r_1$. The first monomer---here marked by \protect\tikz{\protect \draw[draw=red,fill=red] (0,0) circle (0.1cm);}---thus stands for $\mathcal M_1$ (three in this example) different chains of zero length. % \newline (b) Each of the $\mathcal M_1$ chains is extended by one monomer. There are now $M_2=3$ independent chains of length one. Up to now, there is no energy term as there is no bending angle between neighboring bonds. % \newline (c) Each of the $M_2$ chains is extended by one monomer. There are now $M_3=3$ independent chains of length two. % \newline (d) Now, energy comes into play as there is a bending angle between the first and second bond of the polymers. Temperature $T$ and coupling constant $J$ are chosen such that they yield the weights that are given in the sketch ($\times 3,\times 1,\times 0$). Each of the chains is replicated according to its weight. Accordingly $M_{3_{\rm new}}=4$. There are now four independent chains of length two. % \newline (e) Each of these chains is extended independently by one monomer and bond. This procedure is iterated until the desired degree of polymerization is reached. }

\caption{\label{fig:lowDensityRegime}(Color online) End-to-end distribution (a) and tangent-tangent correlations (b) in the low-density regime. The persistence lengths include \mbox{$\xi=0,0.1,0.2,0.3,0.5,0.7,1$} (indicated by the arrow). The occupation probabilities are $p=0$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackSolid.pdf}, black), $0.13$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.25$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.38$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $0.51$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black).}

\caption{\label{fig:endToEndHigh}(Color online) End-to-end distribution function (top) and tangent correlation function (bottom) for $\xi=0.1$ (a), 0.3 (b), and 1 (c). The occupation probabilities are: $p=0$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackSolid.pdf}, black), $p=0.64$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.76$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red), $0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $1$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black). The vertical lines in the end-to-end distribution functions correspond to the distances shown in \fref{fig:gridDistances}.}

\caption{\label{fig:impactDiskDiameter}(Color online) End-to-end distribution for $\sigma=a$ (a) and $\sigma>a$ (b). The occupation probabilities are $p=0$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackSolid.pdf}, black), $0.64$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $0.76$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red),\\* $0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blueDotted.pdf}, blue), and $1$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{blackDashDotted.pdf}, black).}

\caption{(Color online) End-to-end distribution function for $\xi=0.3$ (a) and $\xi=1$ (b). The occupation probabilities are: $p=0.89$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{greenLongDashed.pdf}, green), $1.00$ (\hspace{-0.2cm}\protect\includegraphics[scale=1]{redDashed.pdf}, red). The data marked by $+$ are for chains that are free to move throughout space (no pinpoint) and are done by a multicanonical Monte Carlo simulation. The data marked by $\circ$ are for fixed starting point and are obtained with the growth method.}

125 — 1211.4602

\caption{Density of eigentimes of chaotic cavities with $N$ sites and $M$ leads at the spectrum edge. The scattering matrices belong to circular ensembles CE with the different symmetries ($\beta=1$, $\beta=2$ and $\beta=4$). {\Large $\star$} corresponds to presence of time reversal symmetry (TR) ie $\beta=1$ with $M=2$, $N=23$). {\color{red}$\blacklozenge$} is obtained in absence of (TRS) ie $\beta=2$ (with $M=9$, $N=15$). {\large{\color{darkgreen}$\bullet$}} concerns broken spin rotational symmetry ie $\beta=4$ with ($M=5$, $N=10$). The results are obtained by sampling $\mathcal{H}$ from Lorentzian ensembles with the corresponding $\beta$-symmetry. Fermi energy $E_F=-1.999t_0$.}

126 — 1211.4748

\caption{(color online) Ground-state energies for \elem{O}{16,24}, \elem{Ca}{40,48} and \elem{Ni}{56} as function of $e_{\max}$ for the two types of Hamiltonians (see column headings) using CCSD3B (solid lines) and the NO2B approximation (dashed lines) for a range of flow parameters: $\alpha=0.02\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.04\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.08\,\text{fm}^4$ (\symboltriangle[FGGreen]). }

\caption{ (color online) Ground-state energies for \elem{O}{24}, \elem{Ca}{40,48} and \elem{Ni}{56} as function of $\hbar\Omega$ at $e_{\max}=12$ for the NN+3N-full Hamiltonian using CCSD3B (solid lines) and the NO2B approximation (dashed lines) for a range of flow parameters: $\alpha=0.02\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.04\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.08\,\text{fm}^4$ (\symboltriangle[FGGreen]). }

\caption{ (color online) $\Lambda$-CCSD(T) (solid lines) and CCSD (dashed lines) ground-state energies for \elem{O}{16,24}, \elem{Ca}{40,48} and \elem{Ni}{56} as function of $e_{\max}$ for the two types of Hamiltonians (see column headings) using the NO2B approximation for flow parameters: $\alpha=0.02\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.04\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.08\,\text{fm}^4$ (\symboltriangle[FGGreen]). }

127 — 1211.5145

\caption{\coloronline (a) Dispersion relation before (dashed) and after (solid) a quantum quench opening a gap $\Delta$ and involving only the low-frequency modes. (b) Physical realization: sudden coupling of two independent condensates via a uniform tunneling $j_\perp$. (c) Lattice model used for the numerical calculations.}

\caption{\coloronline Time evolution of the interference contrast $C_{12}(t)=\av{\psi_1\yd\psi_2+{\rm H.c.}}/2N = \av{\cos(\sqrt{2}\phi)}$, according to the two-mode quantum Hamiltonian (\ref{eq:HMF}) with $N=1000$. (a) Exact diagonalization for different values of the ratio $j_\perp/\mu$ = 0.10 {\rm (red)},~0.060 {\rm (green)},0.037 {\rm (blue)},~0.022 {\rm (cyan)},~0.0135 {\rm (magenta)}. (b) Same plots on rescaled axes, showing the universal behavior (\ref{eq:scalinglaw}) with $\eta=0$, superimposed to the analytic solution of the simple pendulum (thick solid line).}

\caption{\coloronline Series expansion (\ref{eq:series}). Numerical results obtained by keeping the first $N_{\rm max}$ terms of the series (solid curve and error-bars) and analytic results \citeSM~for the first-order contribution $N=1$ (dashed curve). (a) $K=1.2,~N_{\rm max}=11$; (b) $K=5,~N_{\rm max}=23$. In both cases we used $10^6$ random evaluation points of $\{(x_i,~t_i)\}_{i=1}^N$ .}

\caption{\coloronline %Time evolution of the interference contrast $C_{12}=\cos(\sqrt{2}\phi)$ obtained from the Truncated Wigner approach (\ref{eq:semiclassic}). (a) Numerical results for $j_\perp/\mu$= 0.27 {\rm (red)},~0.16 {\rm (green)},~0.10 {\rm (blue)},~0.060 {\rm (cyan)},~0.036 {\rm (magenta)} with $K=5$, system size $L=400$, number of random paths $M=400$. (b) Data-collapse according to Eq.~(\ref{eq:scalinglaw2}), superimposed to the result from the series expansion (black curve, identical to Fig.~\ref{fig:montecarlo}(b)).}

\caption{\coloronline %Time evolution of the interference contrast $C_{12}=\av{\cos(\sqrt{2}\phi)}$ obtained from the Hard-core bosons model (\ref{eq:Htebd}). (a) Numerical results for $j_\perp/j_\parallel$= 0.27 {\rm (red)},~0.16 {\rm (green)},0.10 {\rm (blue)},~0.060 {\rm (red)},~0.036 {\rm (cyan)},~0.022 {\rm (magenta)}, with $V/j_\parallel=-1/2$, or $K = \pi/(2\pi-2\arccos(V/2j_\parallel )) \approx 1.2$. The calculations were performed using the Time Evolving Block Decimation (TEBD) algorithm of Refs.~\cite{vidal93,openTEBD} for system size $L=100$, number of states $\chi=100$ and took around three days/curve on a single core. (b) Data collapse according to Eq.~(\ref{eq:scalinglaw2}), superimposed to the result from the series expansion (black curve, identical to Fig.~\ref{fig:montecarlo}(a) after the appropriate rescaling)%to account for the mapping from the microscopic model (\ref{eq:Htebd}) to the effective field theory (\ref{eq:HSG})) .}

\caption{\coloronline Comparison between the idealized case considered in Fig.~2 %\ref{fig:meanfield} and (a) finite initial temperature $T_0$, or (b) finite ramp time $\tau$. The solid lines represent $\delta C_{12}(t) \equiv C^{\rm finite}_{12}(t)-C^{\rm ideal}_{12}(t)$ at different times $t=10/\mu$ (red), $20/\mu$ (green), and $30/\mu$ (blue). The vertical dashed lines correspond to (a) $T_0=\D$, (b) $\tau=1/\D$. As predicted, the dynamics on the left side of these lines is independent of $T_0$ and $\tau$. {\it Numerical parameters --} $j_\perp/\mu=0.01$, ~ $\D=2\sqrt{j_\perp/\mu}=0.2\mu$, and $N=1000$.}

\caption{\coloronline Full distribution function of the spatially integrated coherence (\ref{eq:Z}) , according to the semiclassical equations of motion of (a) the two-mode model and (b) the one dimensional model (11) %(\ref{eq:semiclassic}) . Both subplots represent in the first column a colorplot of the distribution function of $Z$ on the imaginary plane. In the second column, the distribution function of the phase of $Z$. In the third column, the distribution function of the absolute value square of $Z$. {\it Numerical values --} (a) $j_\perp/\mu=0.1$, time is given in units of the inverse plasma frequency $\D=2(j_\perp \mu)^{-1/2}$. (b) $j_\perp/\mu=0.1,~K=25,~L=400$, corresponding to the scaling limit $j_\perp/\mu\ll1$,~$K\gg1$,~$\D L \gg 1$. Time is given in units of the inverse gap $\D^{-1}\approx \Big((4\pi/K) j_\perp \rho_0 \Big)^{-1/2}$.}

128 — 1211.5997

\caption{(a) Effective collection areas after Paris-MVA and energy-dependent point source $\theta^2$ cut; (b) Angular resolution (68\% containment radius) as function of Energy, for point source {\gray}s after Paris-MVA cuts, compared with the standard analysis. (c) Differential sensitivity for a point-source as calculated for array ``I'', for Paris-MVA vs.\standard analysis.}

129 — 1211.6782

\caption{\small \FL\TS map obtained with a maximum likelihood analysis using\emph{front} events above 1\,GeV in the vicinity of\CB. Linear green contours of the 843\,MHz radio map are overlaid~\citep{mcadam1991}. \textcolor{\JKC}{\MOD The blue ellipse compatible with the radio lobes is also shown.} Black cross and contours represent the estimated position and the positional errors at 68\% and 95\% confidence levels in this analysis, respectively. Magenta cross and ellipses represent the position and the errors at 68\% and 95\% confidence levels of 2FGL~J1346.6$-$6027. The position of the core of \CB\is represented as a diamond.\label{fig: cmap_close}}

\caption{\small \SZ/XIS spectrum of the \CB\core. XIS0 and XIS3 spectra are shown in black and red, respectively.\label{fig:core_spe}}

\caption{\small \SZ/XIS FI spectrum of the \CB\lobe region. The observed spectrum are shown in black. Green, red and blue spectra denote the background components: the diffuse X-ray background, the NXB, and the contaminant from the core, respectively.\label{fig:plot_all}}

\caption{\small Broad-band SEDs of \CB\and the model curves for the emission of the extended lobes and core in the system. Red circles denote the total radio fluxes within the range$30$\,MHz\,$-5$\,GHz~\citep[and references therein]{JLM01} and at $30$\,GHz measured by\emph{Planck}. The radio fluxes for the unresolved core, plotted as blue stars, are taken from \citet{JLM01}, \citet{Fey04} and \citet{burke2009}. The \emph{ASCA} measurements of the lobes' and core spectra within the range $0.7-10$\,keV, as discussed in\citet{tashiro98}, are denoted by red and blue bow-ties, respectively. The \emph{Chandra} $0.5-7$\,keV spectrum of the unresolved core is given by cyan bow-tie following\citet{marshall05}. The \FL\fluxes represented by magenta squares correspond to the analysis presented in\S~\ref{sec: obs} of this paper. The black bow-tie denotes the $2-10$\,keV emission of the nucleus emerging from the analysis of our\SZ\data, while the thick black line shows the$2-10$\,keV\SZ\upper limit foe the lobes' emission (see\S~\ref{sec: xray}). Thick gray curves denote the broad-band emission of the lobes modeled in in the framework of the synchrotron and inverse-Compton scenario, while black dashed curve corresponds to the SSC model of the unresolved core in the system, as discussed in \S~\ref{sec: discussion}. \label{fig: model}}

130 — 1212.0227

\caption{Covariant bilinears forming irreps of the lattice symmetry group. Indices $\mu$, $\nu$ and $\rho$ are summed from $1-4$, except that all are different. Pseudoscalar and axial bilinears are not listed: they can be obtained from scalar and vector, respectively, by multiplication by $\gamma_5\otimes\xi_5$. Bilinears related in this way have the same matching factors. This operation also implies the identity of the Z-factors for the tensor bilinears within square brackets. Bilinears marked in \textcolor{blue}{blue} are used as the denominators of the ratios discussed in the text.}

131 — 1212.0402

\caption{101 actions included in UCF101 shown with one sample frame. The color of frame borders specifies to which action type they belong: {\color{NavyBlue} Human-Object Interaction}, {\color{red} Body-Motion Only}, {\color{RoyalPurple} Human-Human Interaction}, {\color{YellowOrange} Playing Musical Instruments}, {\color{ForestGreen} Sports}.}

132 — 1212.0485

\caption{(Color online) (a) DC magnetization ($M$) and its derivative (d$M$/d$H$) for $T =$ 1.4 K. (b) AC susceptibility as a function of magnetic field $H$ for four temperatures. The asterisks and the triangles indicate A-C and C-D phase boundary respectively. (c) $H$-$T$ phase diagram for Ba$_3$NiNb$_2$O$_9$. Different symbols denote phase boundaries obtained from different techniques. $\Box$: $\varepsilon^\prime$ $H$-sweep, {\color{red} $\bullet$}: $\varepsilon^\prime$ $T$-sweep, {\color{red} $\diamond$}: polarization, $\lhd$: $C$$_{\text{P}}$, $\blacktriangle$: AC susceptibility $H$-sweep, {\color{red} $\triangle$}: AC susceptibility $T$-sweep, {\color{blue}$\oplus$}: DC susceptibility $T_{\text{N1}}$, {\color{blue}+}: DC susceptibility $T_{\text{N2}}$, $\ast$: DC magnetization. The arrows indicate spin structures.The dashed (solid) vertical lines in (a) and (b) indicate the transition fields to $uud$ (oblique) phases. The dashed lines in (c) are guides to the eyes.}

133 — 1212.0739

\caption{Energy spectrum of the cosmic-ray iron group. Experimental data as in \fref{iron}. The lines indicate spectra for models explaining the knee due to the maximum energy attained during the acceleration process according to Sveshnikova et al. \cite{sveshnikova} (\line), Berezhko et al. \cite{berezhko} (\dashed), Stanev et al. \cite{stanev} (\dotted), and Kobayakawa et al. \cite{kobayakawa} (\dashdot).\hh}

\caption{Energy spectrum of the cosmic-ray iron group. Experimental data as in \fref{iron}. The lines indicate spectra for models explaining the knee as an effect of the propagation process according to H\"orandel\& Kalmykov et al.\cite{prop} (\line), Ogio et al. \cite{ogio} (\dashed), Roulet et al. \cite{roulet} (\dotted), as well as V\"olk et al.\cite{voelk} (\dashdot).\hh}

134 — 1212.0948

\caption{Dilatometer with thermally decoupled sample. (1)~movable frame, (2)~flat springs, (3)~fixed frame, (4)~sample, (5)~capacitor plates, (6)~sample holder and (7)~thermal insulator, both made from graphite, (8)~thermal anchor, (9)~removable fastening device, (10)~adjustment screw, (11)~piston, (12)~locking screw. Top left: sample holder for a rotation of $\Phi=15^{\circ}$ around the vertical axis. Top right: sketch of the experimental configuration for $ \Delta L \parallel B_\text{in-plane}$ ({\color{red}red}) and $\Delta L \perp B_\text{in-plane}$ ({\color{blue}blue}).}

\caption{Dilatometer with thermally decoupled sample. (1)~movable frame, (2)~flat springs, (3)~fixed frame, (4)~sample, (5)~capacitor plates, (6)~sample holder and (7)~thermal insulator, both made from graphite, (8)~thermal anchor, (9)~removable fastening device, (10)~adjustment screw, (11)~piston, (12)~locking screw. Top left: sample holder for a rotation of $\Phi=15^{\circ}$ around the vertical axis. Top right: sketch of the experimental configuration for $ \Delta L \parallel B_\text{in-plane}$ ({\color{red}red}) and $\Delta L \perp B_\text{in-plane}$ ({\color{blue}blue}).}

135 — 1212.1080

\caption{MEB 3325 actual person positions (\textcolor{green}{O}) and localization estimates (\textcolor{red}{X}) using the forward solving method.}

\caption{MEB 3325 actual person positions (\textcolor{green}{O}) and localization estimates (\textcolor{red}{X}) using the forward solving method..}

136 — 1212.1447

\caption{(color online) Collapse of the PDFs of the CTRW with trapping, for $\alpha=1.5$ and field $\epsilon=0.15$ according to Eq.~(\ref{scaling}) with velocity $v_\epsilon=0.1$. Symbols correspond to different times: $t=10^3$ ($\bullet$), $t=2\cdot 10^3$ ({\color{red}$\blacksquare$}), $t=5\cdot 10^3$ ({\color{green}$\blacklozenge$}), and $t=10^4$ ({\color{blue}$\blacktriangle$}). The continuous line represents the numerical inverse Fourier transform of Eq.~(\ref{four_tr}). Central inset: same data of the main figure in log-log scale, as function of $|x-v_\epsilon t|$. Notice the power law behavior $P_\epsilon(x,t)\sim x^{-(1+\alpha)}$. Top left inset: collapse of the PDF for $\alpha=1.5$ with zero drift, $\epsilon=0$. Notice the simple Gaussian behavior, $\ell(t)\sim t^{1/2}$, in agreement with Eq.~(\ref{four_tr_ep0}).}

\caption{(color online) Collapse of the PDFs with the scaling $t^{2/\alpha} P_\epsilon[(x-v_\epsilon t)/t^{2/\alpha}]$, for the L\'evy walk with exponent$\alpha=2.5$ and field $\epsilon=0.25$, with $v_\epsilon=0.38$. Symbols correspond to different times: $t=1.6\cdot 10^4$ ($\bullet$), $t=3.2\cdot 10^4$ ({\color{red}$\blacksquare$}), $t=6.4\cdot 10^4$ ({\color{green}$\blacklozenge$}), and $t=1.28\cdot 10^5$ ({\color{blue}$\blacktriangle$}). The line represents the numerical inverse Fourier transform of Eq.~(\ref{four_levy}). Inset: same data of the main figure in log-log scale.}

137 — 1212.1708

\caption{\color{black!100}{\sf Ratio of Total decay width and various Branching ratios of the SM-like Higgs boson in the PQ-violating mSuGra to the conserved mSuGra case for $m_0 = 1000, m_{1/2} = 500, A_0 = 0$ (all in GeV units), $\tan\beta = 30, sgn(\mu) = +$.}}

\caption{\color{black!100}{\sf Various Higgs-Boson observables, $R_{ggXY} = \frac{\sigma_{pp\to h} \times{\Gamma}^{h\to X Y}}{\Gamma_h}|_{\frac{SUSY}{SM}}$ with respect to $m_h$ at the LHC.}}

\caption{\color{black!100}{\sf Various Higgs-Boson observables, $R_{ggXY} = \frac{\sigma_{pp\to h} \times{\Gamma}^{h\to X Y}}{\Gamma_h}|_{\frac{SUSY}{SM}}$ with respect to $\lambda_{5}$ at the LHC.}}

138 — 1212.1741

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} Distribution of instantaneous projection coefficient $c_\omega(t)$ for (a) the lowest-frequency mode, mode 0; and (b) the tenth lowest-frequency mode. Red lines are Gaussian fits to the distributions. }

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} Phonon modes in 2D colloidal crystals.~(a) Accumulated mode number, $N(\omega)$, as a function of frequency, for a nearly perfect crystal (blue squares) and an imperfect one (red square); $\omega^2$ is drawn for comparison (black line); small arrows point to the modes whose real space vector distributions are plotted in panels b and c.~(b, c) Spatial distribution of a low-frequency mode for the (b) nearly perfect and (c) imperfect crystal; the direction and magnitude of polarization vectors are represented by the direction and size of the arrows.}

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)}(a) Phonon density of states of two dimensional colloidal crystals. The DOS is obtained from experimental data (black circles), after N/T extrapolation(red squares). The numerically generated DOS for the harmonic triangular lattice (with matched sound speeds) is plotted as a guide to the eye (solid lines). (b) Dispersion curves for longitudinal (open symbols) and transverse (filled symbols) modes along high-symmetry directions, including uncorrected experimental data (black circles), and data after N/T extrapolation (red squares). Theoretical expectations are plotted in dashed (longitudinal) and solid (transverse) lines with matched colors. Inset: high symmetry directions in reciprocal space.}

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} Correction for a finite number of observation frames. (a). Linear dependence of eigenfrequencies on $N/T$ from simulation. $n$ is the mode index which increases from low to high frequencies. For better visualization, a constant, the corresponding mode frequency for $R = 0.002$, has been subtracted for each curve. (b). Linear convergence of eigenfrequencies from experimental data with $N/T$. The vertical axis plots the frequencies minus the frequency from the full length of the video, $\omega_{full}$. (c). Density of states for different ratios of $N/T$ from simulation.}

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} Correction for limited field of veiw. (a). Dispersion curves along different directions derived from uncorrected data. Circles represent longitudinal modes and crosses represent transverse modes. Directions are indexed using the basis vectors of the reciprocal lattice. (b). Dispersion curves along different directions after the Hann window function correction. }

\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} Low-frequency modes in a colloidal crystal with defects.~(a). A snapshot of an imperfect PNIPAM crystal with a grain boundary in the middle of the field of view; the scale bar is 10 \textmu m. (b). Participation ratio for eigenmodes in crystal with defects. (c). Color contour plots indicating polarization magnitudes for each particle, summed over the low-frequency modes with participation ratios less than 0.2. Circles indicate ``defect" particles identified by local structural parameters.}

139 — 1212.2270

\caption{(a) Genuine CV tripartite entanglement for beams $\{1,2,3\}$ can be generated via squeezed states and beam splitters (BS) \cite{aokicv}. (b) Genuine $ $tripartite entanglement exists for $\{1,2',3'\}$ when $Ent_{3}<1$ whereas EPR steering of $\{1\}$ by $\{2',3'\}$ is confirmed when $EPR_{3}=\Delta_{inf,A'}x_{1}\Delta_{inf,A'}p_{1}<1$. Here, $\eta$ is the efficiency of transmission of the eavesdropping beam splitters. Cases $N=2,4$ are also shown. The loss of steering reveals the eavesdropper. \textcolor{red}{}\label{fig:EPRsteering} }

140 — 1212.2323

\caption[aaa]{Modal frequencies (abscissa) and damping factors (ordinates). ({\color[rgb]{1,0,0}{$\circ$}})~: estimations at point~\textbf{A}$_\mathbf{2}$ of the soundboard. Bibliographical results: {{$\diamond$}}~\cite{SUZ1986}; ({\color[rgb]{0,0.5,0}{$+$}})~\cite{DER1997}; ({\color[rgb]{0,0,1}{$\bullet$}})~\cite{BER2003}. --$\,\cdot\,$--~: constant loss-factor curves $\eta~$=\,1, 2 and 3\%.}

\caption[aaa]{Estimations of the modal density in a piano soundboard. Each estimation is the reciprocal of the moving average on six successive modal spacings.\CR Top frame: experimental determinations (apparent modal density) and numerical simulations with mediocre spruce (see below). Modes are measured up to 3~kHz. Dots: 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.~\ref{fig:table_exp_maillage}. {\color[rgb]{.5,.5,.5}$\bm{+}$\hdashrule[0.5ex]{2em}{1pt}{}$\bm{+}$\hdashrule[0.5ex]{2em}{1pt}{}$\bm{+}$}~: numerical modes given by FEM with the wood characteristics of mediocre spruce.\CR Bottom frame: numerical modes given by FEM with the wood characteristics of Norway spruce ({\color[rgb]{0,0,1}$\bm{\Box}$}), Sitka spruce ({\color[rgb]{1,0,0}$\bm{\times}$}) and mediocre spruce ({\color[rgb]{.5,.5,.5}$\bm{+}$\hdashrule[0.5ex]{2em}{1pt}{}$\bm{+}$\hdashrule[0.5ex]{2em}{1pt}{}$\bm{+}$}). }

\caption[aaa]{Typical bank-filtering analysis of a reconstructed impulse response between 550 and 1150~Hz (acoustical excitation). --~--~--\enskip,\enskip$\cdots$\enskip,\enskip--\,$\cdot$\,--\enskip:~amplitude responses of the (slightly overlapping) narrow-band filters. {\color[rgb]{0,0,1}------}\,: Fourier spectrum of the impulse response at point\textbf{A}$_\mathbf{2}$. {\color[rgb]{1,0,0}$\bullet$\,}: modes estimated by the high-resolution modal analysis (modal amplitudes and frequencies).}

\caption[aaa]{Modal frequencies and dampings in the \mbox{[0-600] Hz} frequency-band, obtained after an \emph{impulse} excitation and given by a high-resolution modal analysis. (a)~Direct analysis. (b)~Analysis including a narrow band-pass filtering. (c)~Results after suppression of the (usually poor) estimations in the nodal regions. ($\circ$)~: retained modal parameters. ({\color[rgb]{0,0.5,0}$\times$})~: weighted means of the modal parameters estimated at four points of the soundboard with an \emph{acoustical} excitation (see following section). --$\,\cdot\,$--~: constant loss-factors $\eta~$=\,1, 2 and 3\%.}

141 — 1212.2982

\caption{Simulations illustrating the effect of mixture on $X^2$ in two-qubit tomography. (a--c) Werner states~\cite{WernerRF1989} of the form $\rho_t = p\ket{\psi}\bra{\psi} + (1-p) I_4/4$ ($I_4$ is the identity operator in four dimensions and $\ket{\psi}$ is a maximally entangled state) studied as a function of $p$. For each value of $p$, randomly chosen states are used to generate Poissonian-distributed noisy data sets for different platonic solid measurement sets~\cite{deBurghMD2008}: tetrahedron (minimally complete, red: {\color[rgb]{1,0,0}---$\circ$---\color[rgb]{1,0,0}}), cube (over-complete, blue: {\color[rgb]{0,0,1}---$\Diamond$---\color[rgb]{0,0,1}}), octahedron (green: {\color[rgb]{0,1,0}---$\Box$---\color[rgb]{0,1,0}}), and dodecahedron (magenta: {\color[rgb]{1,0,1}---$\times$---\color[rgb]{1,0,1}}); which are then reconstructed using maximum likelihood estimation. Each point represents an average over 500 random states, except in the four-measurement set data, which averages over 7500 states per point. The graphs show (a) the average chi-squared test statistic for the noisy reconstructed states (the much-larger-valued twelve-measurement set exhibits the same trend, but is omitted in order to show better detail with the other sets); and (b) the fraction of reconstructed states which are full rank; and (c) the average fidelity of the noisy reconstructed states with the known target state. In (a), the solid lines show the expected value of the test statistic, assuming $M-d^2$ DOFs and the dashed lines show the expected position of the first standard deviation higher. The dotted lines show the width of the observed distributions (because they are skewed, the lines show the $\chi^2$-function positions which correspond to the integrated probabilities for the $\pm \sigma$ positions of a symmetric Gaussian distribution). (d--e) 150,000 randomly chosen, mixed two-qubit state reconstructions (6 measurements per qubit) with different numbers of nonzero eigenvalues. (d) Comparison of the observed distribution of $X^2$ with the theoretical $\chi^2$ probability distribution for $M-d^2=20$ DOFs (red curve) and an optimal $\chi^2$ distributions with a fit to the number of DOFs over arbitrary real numbers (green curve: $M{-}c \sim 23.97$). Vertical lines show the expected means and standard deviation positions. (e) $X^2$ vs purity ($\Tr{\varrho^2}$) of the reconstructed states. Dotted lines at $\Tr{\varrho^2}=1/4$, 1/3 and 1/2 show the minimal purities for states with 4, 3 and 2 nonzero eigenvalues, respectively.}

142 — 1212.3119

\caption{(Left) Evolution of SDR as a function of CPU time (in seconds), for (\textcolor{blue}{blue}) our method and \textcolor{red}{(red)} NMF started from several initial points. (Right) Zoom on the first few seconds.}

143 — 1212.3339

\caption{ The redshift sensitivity \blue{of the modified gravity parameter $\mu_0$}. The weight function $\phi(z)$ is defined in equation (\ref{eq:wt_fn}) \blue{and evaluated with equation (\ref{eq:pca})}. \label{fig:weight_function}}

\caption{\label{tab:results} Parameter constraints for different combinations of dataset and parameter space. The three different backgrounds we explore are flat \lcdm, flat \wcdm and the non-flat \wcdm which we denote as o\wcdm. \blue{All constraints make use of the small scale anisotropies from WMAP7 ($\ell \geq 100$), while those marked ISW also utilise the larger angular scales.} }

\caption{The cosmic shear data from CFHTLenS. \red{Each panel contains the two lensing correlation functions, $\xi_+(\theta)$ (filled) and $\xi_-(\theta)$ (empty) with the $\xi_-(\theta)$ points slightly offset for clarity. The source galaxies are divided into two redshift bins, a `low' redshift bin with $0.5<z \leq 0.85$ and a `high' redshift bin with $0.85<z \leq 1.3$. The left, middle and right hand panels display data from the low-low, low-high, and high-high correlations. }In each case, the black lines are the set of theoretical curves corresponding to $\xi_+(\theta)$ (upper) and $\xi_-(\theta)$ (lower) for the best-fitting \lcdm cosmology. The dashed lines correspond to a cosmological model with the same expansion history $H(z)$ as the best-fitting model, but with a boost in the effective gravitational constant ($\mu_0 = 0.5; \Sigma_0 = 0$). The dot-dashed lines demonstrate the signal one expects if the present-day ($z=0$) gravitational potentials are associated with Newtonian gravity ($\mu_0 = 0; \Sigma_0 = -0.5$).\label{fig:cfht_data}}

\caption{The growth data from WiggleZ \blue{\citep{Blake2012MNRAS}} and 6dFGS \citep{Beutler2012MNRAS}. The black line is the corresponding theoretical curve for the best-fit \lcdm cosmology. The dashed lines correspond to the same cosmology but with a boost in effective gravitational constant $\mu_0$. \label{fig:fsig8_data}}

144 — 1212.3496

\caption{ Profile of the three dimensional MHD blast wave test without AMR using 1 M cells. Allreduce labels the only global communication executed each time step where the maximum length of the physical time step is obtained using {\tt MPI\_Allreduce}. Initialization and file I/O are not included. }

\caption{ Profile of the three dimensional MHD blast wave test with AMR using at most 10 M cells. Allreduce labels the calculation of the maximum length of the physical time step obtained using {\tt MPI\_Allreduce}. Initialization and file I/O are not included. }

145 — 1212.4323

\caption{\label{fig:three}(Color on-line) a) The imaginary part {\em vs} the real part of the complex rotation eigenvalues . All the eigenvalues were calculated for $B=5T$ and different rotation angles. From top to bottom, the eigenvalues correspond to $\theta=0.1$ (black \fullcircle), $\theta=0.11$ (red {\color{red}\fullsquare}), $\theta=0.12$ (green {\color{green} $\blacklozenge$}), up to $\theta=0.2$. The three fan-like sets of data are related to the first three LL's. Is is clear that the the leftmost fan and the central one start to overlap around $27 \, \mbox{meV}$. For $B<5T$ the overlap grows larger making more difficult to look for the resonance data. b) The $\theta$ trajectory of the resonance eigenvalue. The symbols correspond to the angles shown in panel a). The eigenvalues that form the $\theta$ trajectory can also be observed in panel a), they lie near the bump that presents the leftmost fan around $2 \, \mbox{meV}$ }

\caption{\label{fig:five} The function $\mathcal{G}_n$ {\em vs} the magnetic field. The figure shows the function for the first six variational eigenvalues, $\mathcal{G}_1$ , $\mathcal{G}_2$, $\ldots$, $\mathcal{G}_6$, solid (\full), dotted (\dotted), short-dashed (\dashed), long-dashed (\longbroken), dot-dashed (\chain) and double dot-dashed (\dashddot) lines respectively. The peak where each curve attains its maximum value is clearly appreciable. $B_n$ is given by the abscissa of the peak (see the text). }

\caption{\label{fig:six}(Color on-line) The critical fields $B_p(a_{\rho})$ and $B_{\mathcal{F}}(a_{\rho})$ {\em vs} the quantum dot radius. The localization probability critical field (red squared dots, {\color{red}\fullsquare}) and the fidelity critical field (solid black dots, \fullcircle) data is shown for several quantum dot radius. The (blue) dashed curve correspond to the radius of the lowest LL as a function of the magnetic field. It is clear that for small quantum dot radius the localization takes place when the cylindrical wave function enters into the cylindrical QD. For larger QD's radius the localization is dominated by the quantum well potential and not by the magnetic field. }

\caption{\label{fig:seven} The resonance energy $E_{res}$ calculated using the three methods depicted in the text, complex exterior scaling, localization probability, fidelity, the corresponding values are shown using blue square dots ({\color{blue}\fullsquare}), black circles (\opencircle), and magenta diamonds ({\color{magenta} \opendiamond}). The energy of the resonance in the region above the second LL can be obtained using a modified version of the complex rotation method and is shown using blue open squares ({\color{blue} \opensquare}) . }

146 — 1212.4947

\caption{$\alpha(n_0)$ and $P(n_0)$ in the Debye-H\"uckel model at$T=20\,000\ \mathrm K$ (\full), $T=23\,000\ \mathrm K$ (\broken), and $T=26\,000\ \mathrm K$ (\dashed).}

\caption{$\alpha(n_0)$ and $P(n_0)$ in perturbed Debye-H\"uckel plasmas at$T=23\,000\ \mathrm K$ for\\ (a)--(b): $a=0$ (\full), $a\,T/e^4\approx-0.04$ (\broken), and $a\,T/e^4\approx0.04$ (\dashed), $b,\,c=0;$\\ (c)--(d): $b=0$ (\full), $b\,T/e^2\approx-0.06.$ (\broken), and $b\,T/e^2\approx0.06.$ (\dashed), $a,\,c=0;$\\ (e)--(f): $c=0$ (\full), $c\,e^2\,T\approx0.035$ (\broken), and $c\,e^2\,T\approx0.07.$ (\dashed), $a,\,b=0.$}

147 — 1212.6640

\caption{Oracle\textregistered\RDBMS architecture}