2010

1 — 1001.2428

\caption{Plot of $I/c$ as a function of $\Delta \Delta G$ for four different experiments at concentrations $c=2$, $10$, $50$ and $250$ pM. The hybridization time is of 17 h. \textcolor{Black}{The ``collapse" of the four data sets into a single curve demonstrates that $I \propto c$ in the whole intensity range.} Deviations from the equilibrium isotherm $I \propto e^{-\Delta G/RT}$ are observed at high intensities. Inset: plot of each individual concentration. }

\caption{Plot of intensity vs. $\Delta \Delta G$ four experiments at different times with a $30$-mer target and $c=50$ pM. Dashed lines have slope $1/RT$, %solid \textcolor{Black}{dotted} lines have slope $\gamma/RT$ with $\gamma = 0.32$. Solid lines are obtained from the solution of Eqs.~(\ref{dtheta1dt}) and (\ref{dtheta2dt}) using % the parameters of Fig.~\ref{fig_timesL30}. \textcolor{Black}{as parameters $k_1 =10^5 M^{-1} s^{-1}$ , $k_2=1s^{-1}$, $\gamma = 0.32$ and $\Delta G_{\rm PM} = -14.5$ kcal/mol}. }

\caption{As in Fig.~\ref{fig_timesL30} for a target sequence of length $25$. \textcolor{Black}{The target concentration is $c=500$pM}.}

2 — 1001.2573

\caption{ Formation energy $E_{f\!orm}$ in eV per TiO${}_{\text{2}}$ functional unit versus TiO${}_{\text{2}}$ structure for 0D (TiO${}_{\text{2}}$)${}_n$ clusters ($n\leq \text{9}$), 1D TiO${}_{\text{2}}$ (2,2) nanorods, (3,3) nanotubes, and (4,4) nanotubes, 2D HexABC and anatase layers, and 3D anatase surface, anatase bulk, and rutile bulk phases. DFT calculations using RPBE for the undoped ({\color{red}{$\medcirc$}}, red), nitrogen doped ({\color{blue}{$\boxempty$}}, blue), and boron doped ({\color{magenta}{$\vartriangle$}}, magenta) systems are shown. Note the $sp^{\text{2}}$ character of oxygen in the more stable rutile and (2,2) nanorod structure, and the $sp^{\text{3}}$ character of oxygen in the less stable layer and larger nanotube structures.}

\caption{ Formation energy $E_{f\!orm}$ in eV per TiO${}_{\text{2}}$ functional unit versus radius $R$ in \AA~for TiO${}_{\text{2}}$ (2,2) nanorods and ($n,n$) nanotubes ($3\leq n \leq 8$) ({\color{red}{$\medcirc$}}, red). For comparison, the TiO${}_{\text{2}}$ nanotube formation energies are also approximated by $E_{f\!orm} \approx E_{layer} + (E_{rutile} - E_{layer}) R_{\text{(2,2)}}/R$, ({\color{red}{\textbf{---$\!$---$\!$---}}}, red), where $R_{\text{(2,2)}} \approx $2.54 \AA~is the radius of a TiO${}_{\text{2}}$ (2,2) nanorod.}

\caption{Energy gap $\varepsilon_{gap}$ in eV versus TiO${}_{\text{2}}$ structure for 0D (TiO${}_{\text{2}}$)${}_n$ clusters ($n\leq \text{9}$), 1D TiO${}_{\text{2}}$ (2,2) nanorods, (3,3) nanotubes, (4,4) nanotubes, 2D HexABC and anatase layers, and 3D anatase surface, anatase bulk, and rutile bulk phases. DFT calculations using RPBE of the highest occupied and lowest unoccupied state gaps / $I_p - E_a$ for the (a) undoped ({\color{red}{$\medcirc$}}/{\color{red}{\medbullet}}, red), (b) boron doped ({\color{magenta}{$\vartriangle$}}/{\color{magenta}{$\blacktriangle$}}, magenta), and (c) nitrogen doped ({\color{blue}{$\boxempty$}}/{\color{blue}{$\blacksquare$}}, blue) systems are compared with undoped B3LYP ({\color{red}{$\blacktriangledown$}}), GW ($\blacktriangleright$), and experimental ($\Diamondblack$) results~\cite{Structures,GW_TiO2_anatase,N-TiO2NTarraydoping,TiO2Rev1}. Small open symbols denote transitions between highest fully occupied states and the conduction band. }

3 — 1001.5233

\caption{(color online). (a) State $j$ mean relative population $N_j/N$ as a function of interrogation time $T$, for (\red{$\circ$}) $j=1$ and (\blue{$\times$}) $j=2$, following an $S_{N}(T)$ sequence, with $\ell=1$. (b) Angular density fringes of state $1$, with $\ell=1$, $m=0$, $\Omega=1$, for ($\diamond$) $T = 0$, ($+$) $T = 0.1$, (\green{$\square$}) $T = 1$, and (\magnta{$\bigtriangleup$}) $T = 8.3$, following an $S_{F}(T)$ sequence. The interaction strengths $g_{jk}$ correspond to an $^{87}$Rb configuration. All quantities are dimensionless.}

4 — 1002.0061

\caption{\label{fig:emq_La} The \qv\dependencies of the Fe magnetic moment (a) and the total energy (b) calculated for\LFAO\with$\delta$=0 (red \fullcircle), 0.1 (magenta $\blacktriangledown$), and 0.2 (green $\blacksquare$).}

\caption{\label{fig:emq_BaSr} The \qv\dependencies of the Fe magnetic moment (a, c) and the total energy (b, d) calculated for\BFA\(a, b) and \SFA\(c, d) with AF alignment of Fe moments along the $c$ axis ($q_z=\pi/d_{\perp}$) and $\delta$=$-$0.1 (blue $\blacktriangle$), 0 (red \fullcircle), and 0.1 (magenta $\blacktriangledown$). The corresponding curves calculated for undoped compounds ($\delta$=0) with $q_z$=0 are also shown (red \opencircle).}

5 — 1002.0258

\caption{(a) Reconstructed IXS phonon spectrum of HgBa$_{2}$CuO$_{4}$ from \cite{Uchi04} for $\vc{q}=\frac{2\pi}{a}(\zeta,0,0)$ using a Lorentz fit with one mode only (thick line), the corresponding Lorentz peaks of the modes are also shown (thin line). The broken line represents the elastic background. The inelastic cross section $N(E)$ is given in arbitrary units. (b) Same as in (a) but using a Lorentz fit on the basis of the calculated phonon dispersion for modes of $\Delta_1$-symmetry in the energy range considered \cite{Bauer10}. (c) Calculated phonon dispersion of HgBa$_{2}$CuO$_{4}$ for modes of $\Delta_1$-symmetry according to \cite{Bauer10}. The symbols \fullsquare\and\opensquare\denote the phonon frequencies from the Lorentz fit and\fullcircle\the half-breathing mode. The vertical bars indicate the FWHM of the peaks determined in the fitting.}

\caption{(a) Reconstructed IXS phonon spectrum of Bi$_{2}$Sr$_{2}$CuO$_{6}$ from \cite{Graf08} for $\vc{q}=\frac{2\pi}{a}(\zeta,0,0)$ using a Lorentz fit with two modes only (thick line), the corresponding Lorentz peaks of the modes are also shown (thin line). The broken line represents the elastic background. The inelastic cross section $N(E)$ is given in arbitrary units. (b) Same as in (a) but using a Lorentz fit on the basis of the calculated phonon dispersion for modes of $\Delta_1$-symmetry in the energy range considered \cite{Bauer10}. (c) Calculated phonon dispersion of Bi$_{2}$Sr$_{2}$Cu$_{6}$ for modes of $\Delta_1$-symmetry according to \cite{Bauer10}. The symbols \fullsquare\and\opensquare\denote the phonon frequencies from the Lorentz fit and\fullcircle\the half-breathing modes. The vertical bars indicate the FWHM of the peaks determined in the fitting.}

6 — 1002.3420

\caption{The simultaneous fit made to the EPIC-pn and RGS spectra of Obs. 3, by using the best fit model of Obs. 2 and fitting only the covering fraction of phase B. The \redchi is 130 (2710 d.o.f.). The RGS data are shown in blue, the EPIC-pn data in red and the models in black.}

\caption{The simultaneous fit made to the EPIC-pn and RGS spectra of Obs. 3, using the best fit model of Obs. 2 and fitting only the continuum parameters (maintaining $\fcov=0.55$). The \redchi is 1.76 (2707 d.o.f.). The RGS data are shown in blue, the EPIC-pn data in red and the models in black. The significance of the \ion{O}{vii} He-$\gamma$ and \ion{Ca}{xiv} lines is discussed in Sect. \ref{three_phase}.}

7 — 1002.3603

\caption{\label{fig:data}LAT measured \gray{} intensity with fit results for $|b| \geq 10^\circ$ including statistical and systematic errors. Fit results by component are given in Table~\ref{tab:table1}. Note LAT data are dominated by systematic uncertainties for the energy range shown in the figure.}

\caption{\label{fig:galcomp}LAT measured \gray{} intensity with fit results for $|b| \geq 10^\circ$ including statistical and systematic errors. Independently fitted components of the galactic diffuse emission are shown individually. ``Other Galactic diffuse'' denotes the sum of the \htwo{} and non-local \hi{} and \hii{} contributions which are included using the model intensities (i.e., no fit). Note LAT data are dominated by systematic uncertainties for the energy range shown in the figure. Ths intensity of the solar emission considered in the model is below the range shown in the figure. }

\caption{\label{fig:latitudes} LAT measured intensity compared to the \gray{} emission model used in the derivation of the EGB averaged over different ranges in Galactic latitude. The regions shown are $10^{\circ} \leq |b| \leq 20^{\circ}$ (upper left), $20^{\circ} \leq |b| \leq 60^{\circ}$ (upper right) and $|b| \geq 60^{\circ}$ (lower center). ``Other Galactic diffuse'' denotes the sum of the \htwo{} and non-local \hi{} and \hii{} contributions which are included using the model intensities (i.e., no fit). Errors include statistical and systematic errors. LAT data are dominated by systematic uncertainties for the energy range displayed in the figure.}

\caption{\label{fig:hemispheres} LAT measured intensity compared to the \gray{} emission model used in the derivation of the EGB averaged over different hemispheres on the sky for Galactic latitudes $|b| \geq 10^{\circ}$. The hemispheres shown are centered at the North Galactic pole (upper left), the South Galactic pole (upper right), the Galactic center (lower left) and anti-center (lower right). ``Other Galactic diffuse'' denotes the sum of the \htwo{} and non-local \hi{} and \hii{} contributions which are included using the model intensities (i.e., no fit). Errors include statistical and systematic errors. LAT data are dominated by systematic uncertainties for the energy range displayed in the figure.}

8 — 1002.4438

\caption{Datasets, papers and the URLs where the datasets can be downloaded. \textcolor{red}{$\odot$} and \textcolor{cyan}{$\blacksquare$} denote which datasets are synthetic and real respectively.}

9 — 1002.4599

\caption{Gaussian kernel density estimate ({\color{blue}---}) of $\Psi(\psi,t)$ for the measured outputs ({\color{blue}o}) and the associated Gaussian kernels ({\color{red}-- --}).}

\caption{Real ({\color{blue}---}) vs. estimated ({\color{red}-o-}) parameter distribution, with grid points ({\color{red}o}).}

10 — 1002.4834

\caption{Formation energy \(\Delta E_f\) of rutile metal oxide (2,2) nanorods ({\footnotesize\medbullet}), (3,3) nanotubes ({\medbullet}), and (110) surfaces ($\blacksquare$) in eV/MO$_{\text{2}}$ vs.~(a) bulk heat of formation \(\Delta G_f^{\text{Bulk}}\)for rutile metal oxides in eV/MO$_2$ from Ref.~\onlinecite{Martinez}, (b) $d$-band center \(\varepsilon_{d}\) for the bulk metal oxide relative to the Fermi energy in eV, and (c) fraction of M--O bonds broken times the $d$-band center \(\Delta \chi \varepsilon_{d}\) in eV/MO$_2$. The experimental nanoparticle surface energy \cite{SurfaceEnthalpiesTiO2} and bulk formation energy \cite{CRCHandbook,Pourbaix} for rutile TiO$_2$ ({\color{red}{$\boxempty$}}) are provided for comparison. }

11 — 1003.0353

\caption{\label{fig2} (colour online) Collapse times (open symbols) and revival times (filled symbols) versus % the inverse many-body interaction strength $1/g$ for $N,L=5,5$ and $V_0=4$: \opensquare, $V_0 = 5$: \opencircle, $V_0 = 6$: \opendiamond. Order of the resonance: $r=1$. The collapse time is defined via $\mathcal N_b(t_{\rm coll}) = (1+\rme^{-1})/2$ and the revival time is chosen as the next maximum in the oscillations after initial decay $t> t_{\rm coll}$, as indicated in \fref{fig1} (upper panel). % A linear fit yields: $t_{\rm rev}/T_B = 718/g$ for $V_0= 4 E_r$ and $t_{\rm rev}/T_B = 700/g$ for $V_0= 5 E_r$. Parameters: $N,L=5,5$. }

\caption{\label{fig3} (colour online) $|c_n|$ versus the corresponding quasi-energies $\varepsilon_n$ in an expansion in the eigenbasis of the Floquet--Bloch operator \eref{eq:FBoperator}. Shown are different interaction strengths: $g=0.0$ (*), $g = 0.05$ (\fullcircle$\!$), $g = 0.1$ (\fullsquare), and $g = 0.2$ (\fulldiamond); other parameters as in \fref{fig1}. We observe that different quasi-energies are shifted at different rates when increasing the interaction strength.}

\caption{\label{fig4} (colour online) Comparison between estimated revival times according to \eref{eq:revivaltime} (\broken) and to \eref{eq:revivaltime2} (\fullcircle$\!\!$) with numerical simulations (\opensquare) for $V_0=8, r=1$, and different system sizes, parametrised by the Hilbert space dimension of the $\kappa=0$ subspaces. The error bars indicate the width of the revival at half maximum. \emph{Inset:} Scaling of the numerically measured revival time with the lattice depth $V_0$ for $r=1$. We multiply $t_{\rm rev}$ by $W_x$ since $W_x$ also depends on $V_0$ to show the remaining non-trivial scaling behaviour. Shown are $N,L=4,4$ ($\dim \mathcal H = 86$, \opensquare), $N,L=5,5$ ($\dim \mathcal H = 402$, \opendiamond), % $N,L=6,6$ ($\dim \mathcal H = 2076$, \opentriangle), $N,L=6,7$ ($\dim \mathcal H = 3876$, \opentriangle) $N,L=7,7$ ($\dim \mathcal H = 11076$, $\times$). Filled symbols are $t_{\rm rev}$ expected from \eref{eq:revivaltime2} for the same system, and the thick dashed line again shows our universal result \eref{eq:revivaltime}.}

12 — 1003.1053

\caption{Fixed points and stable manifolds. The set of fixed points of the stroboscopic map \eqref{strobo}, for $D=1$ and varying period $\tau$, which is made up of three different branches is displayed. As an example the stable (\textcolor{red}{$\bullet$}\textcolor{green}{$\bullet$}\textcolor{blue}{$\bullet$}) and unstable ($\circ$) fixed points corresponding to $\tau=4$ are shown along with the stable manifolds of the unstable fixed points (brown surfaces). T cell concentrations are represented logarithmically. }

13 — 1003.1121

\caption{(color online) (a) Particle fluctuation relaxation time increases abruptly with wall strain amplitude across the entire channel. Inset: Sample mean square displacement time series for $\gamma_0=2.8$ (\textcolor{red}{$\square$}) and $\gamma_0=4.0$ (\textcolor{green}{$\circ$}) at $y/a=0.44$ with fits. (b) The local critical strain, $\gamma^c$, observed at the global onset $\gamma_0^c \approx 3.4$, as a function of the local concentration, $\phi$. The much steeper variation here compared with the uniformly strained case demonstrates that concentration variations alone are insufficient to account for the simultaneous criticality across the channel.}

\caption{(color online) Particle motions in the center of the channel become correlated over a longer range above the critical wall strain amplitude, which indicates that the fluctuations do not arise from local interactions. Inset: Sample correlation length profiles for $\gamma_0=2.3$ (\textcolor{red}{$\square$}) and $\gamma_0=4.0$ (\textcolor{green}{$\circ$}).}

14 — 1003.1346

\caption{\label{Figure:Results1350}(Color online) The photon beam asymmetries extracted from the data set with a coherent peak position at 1305~MeV. The filled (red) circles {\color{red} ($\bullet$)} denote this analysis, the (green) stars {\color{green} ($\ast$)} denote our previous CBELSA/TAPS analysis~\cite{Elsner:2008sn}, and the open (blue) circles {\color{blue} ($\circ$)} denote the GRAAL results~\cite{Bartalini:2005wx}. The black solid line shows the recently published solution of the Bonn-Gatchina partial wave analysis (PWA)~\cite{Anisovich:2009zy}, the gray solid line denotes the SAID SP09 prediction~\cite{said:database,Dugger:2009pn}, the dashed black line shows the recent MAID solution~\cite{Drechsel:2007if}, and the dashed-dotted line shows a new preliminary solution of the Bonn-Gatchina PWA that includes the results of this analysis \cite{Andrey:private}. The width of the energy bins is 33~MeV, consistent with the earlier published results. The energy of the bin centers is given in each distribution.}

\caption{\label{Figure:Results1600}(Color online) The photon beam asymmetries extracted from the data set with a coherent peak position at 1610~MeV. The filled (red) circles {\color{red} ($\bullet$)} denote this analysis, the (green) stars {\color{green} ($\ast$)} denote our previous CBELSA/TAPS analysis~\cite{Elsner:2008sn}, the open (blue) circles {\color{blue} ($\circ$)} denote the GRAAL results~\cite{Bartalini:2005wx}, and the (blue-green) stars {\color{cyan} ($\ast$)} above 1500~MeV denote recent LEPS results~\cite{Sumihama:2007qa}. The black solid line shows the recent solution of the Bonn-Gatchina PWA~\cite{Anisovich:2009zy}, the gray solid line denotes the SAID SP09 prediction~\cite{said:database,Dugger:2009pn}, and the dashed black line shows the recent MAID solution~\cite{Drechsel:2007if}. The width of the energy bins is 33~MeV, consistent with the earlier published results. The energy of the bin centers is given in each distribution. For energies below 1400~MeV, we have averaged the results from both data samples.}

15 — 1003.2341

\caption{Electronic band structure of rutile bulk, along the high symmetry directions of the irreducible Brillouin zone, from GGA calculations ({\textbf{---$\!$---}}). and including the $G_0W_0$ correction ({\color{yellow}{\medbullet}}).}

\caption{Imaginary part of the dielectric constant for rutile (left), and anatase (right), in-plane polarization, calculated by GGA RPA ({\color{blue}{\textbf{-- -- --}}}), using $G_0W_0$ on top of GGA ({\color{red}{$\cdots\cdots$}}), and via the Bethe-Salpeter equation (BSE) ({\color{black}{\textbf{---$\!$---}}}). The experimental spectrum ({---~\textbullet~---}) from Refs.~\cite{Cardona} and \cite{PhysRevB.61.7459} is also shown for comparison.}

\caption{Energy gap $\varepsilon_{gap}$ in~eV versus TiO${}_{\text{2}}$ structure for 0D (TiO${}_{\text{2}}$)${}_n$ clusters ($n\leq {9}$), 1D TiO${}_{\text{2}}$ (2,2) nanorods, (3,3) nanotubes, (4,4) nanotubes, 2D HexABC and anatase layers, and 3D anatase surface, anatase bulk, and rutile bulk phases. DFT calculations using the highest occupied and lowest unoccupied state gaps with the standard GGA RPBE xc-functional ({\color{red}{$\medcirc$}}) and the hybrid B3LYP xc-functional ({\color{red}{$\blacktriangledown$}}), are compared with DFT $I_p - E_a$ calculations ({\color{red}{\medbullet}}), $G_0W_0$ quasi-particle calculations ({\color{red}{$\blacktriangleright$}}), and experimental results (\Diamondblack)~\cite{TiO2,Structures,GW_TiO2_anatase,N-TiO2NTarraydoping,TiO2Rev1}. Schematics of representative structures for each dimensionality are shown above and taken from Ref.~\cite{TiO2}. }

\caption{Total density of states (DOS) in~eV$^{-1}$ per TiO$_{\text{2}}$ functional unit vs.~energy \(\varepsilon\) in~eV for undoped ({\color{black}{---$\!$---}}), boron doped ({\color{magenta}{\textbf{-- -- --}}}) and nitrogen doped ({\color{blue}{\textbf{---$\!$---}}}) (2,2) TiO$_{\text{2}}$ nanorods from standard DFT GGA RPBE xc-functional and (inset) $G_0W_0$ quasi-particle calculations. The highest occupied states for boron and nitrogen doped (2,2) TiO$_{{2}}$ nanorods are depicted by isosurfaces of $\pm$0.05$e/$\AA$^{{3}}$ in the structure diagrams to the left.}

\caption{Influence of doping on the energy gap $\varepsilon_{gap}$ in~eV versus TiO${}_{\text{2}}$ structure for 0D (TiO${}_{\text{2}}$)${}_n$ clusters ($n\leq {9}$), 1D TiO${}_{\text{2}}$ (2,2) nanorods, (3,3) nanotubes, (4,4) nanotubes, 2D HexABC and anatase layers, and 3D anatase surface and bulk. The energy gaps for (a) boron doped systems from DFT calculations using the highest occupied and lowest unoccupied state gaps with the standard GGA RPBE xc-functional ({\color{magenta}{$\vartriangle$}}) are compared with DFT $I_p - E_a$ calculations ({\color{magenta}{$\blacktriangle$}}), and $G_0W_0$ quasi-particle calculations ({\color{magenta}{$\blacktriangleright$}}), and (b) nitrogen doped systems from DFT calculations using the highest occupied and lowest unoccupied state gaps with the standard GGA RPBE xc-functional ({\color{blue}{$\boxempty$}}) are compared with DFT $I_p - E_a$ calculations ({\color{blue}{$\blacksquare$}}), and $G_0W_0$ quasi-particle calculations ({\color{blue}{$\blacktriangleright$}}), and experimental (\Diamondblack) results~\cite{TiO2,Structures,GW_TiO2_anatase,N-TiO2NTarraydoping,TiO2Rev1}. Small open symbols denote transitions between highest fully occupied states and the conduction band.}

16 — 1003.3182

\caption{Simulated behavior of the time-averaged mean squared displacement $\left<\overline{\delta^2(\Delta)}\right>$ for a continuous time random walk process with waiting time distribution $\psi(t)\sim\tau^{\alpha}/t^{ 1+\alpha}$, in an harmonic binding potential $U(x)=x^2$ ($\square$), and in a box with reflecting boundary conditions and size $2$ ($\circ$). The anomalous diffusion exponent is $\alpha = 1/2$, and the measurement time was chosen as $T=10^7$ (a.u.). We also chose $k_BT_{mp}=0.1$, and $K_{1/2}=0.0892$. Without fit, the lines show the analytic results for the transition from the initial linear lag time dependence $\simeq\Delta^1$, Eq.~\eqref{freepart} (--- and $-\cdot-$), to the long lag time behavior $\simeq\Delta^{1-\alpha}$, Eq.~\eqref{msd01} (---). In both cases $\langle x\rangle_B=0$. At long lag times $\left<\overline{\delta^2(\Delta)}\right>/\langle x^2\rangle_B$ exhibits universal behavior, independent of the external field. \label{fig1}} \end{figure} Fig.~\ref{fig1} shows, over a large time span, the time-averaged mean squared displacement $\left<\overline{\delta^2(\Delta)}\right>$ of a subdiffusing particle in (i) an harmonic potential, and (ii) in a box with reflecting boundaries. The initial particle position was chosen to be at the bottom of the potential and in the center of the box, respectively. At short lag times $\Delta$ we observe the linear scaling \begin{equation} \left<\overline{\delta^2(\Delta)}\right>\sim\frac{2K_\alpha\Delta}{\Gamma(1+\alpha)T^{1-\alpha}} \label{freepart} \end{equation} with the lag time $\Delta$. In this result only the dependence on the overall measurement time $T$ bears witness to the fact that the underlying stochastic process is subdiffusive. Seemingly paradox, the lag time $\Delta$ enters linearly, in contrast to the associated ensemble averaged mean squared displacement $\langle x^2(t)\rangle\sim 2K_{\alpha}t^{\alpha}/\Gamma(1+\alpha)$. However, this is the result of the weak ergodicity breaking of the process, as shown in Refs.~\cite{Klafter,He}. The free particle behavior at short $\Delta$ is an expected result, which can be obtained from scaling arguments or explicitly from the full correlation function \cite{unpublished}: at sufficiently short times the particle does not yet feel the confinement due to the reflecting boundaries, or it does not yet experience the restoring force of the potential, respectively. The $\left<\overline{\delta^2(\Delta)} \right>\simeq\Delta$ regime holds for scales of the lag time $\Delta$ that fulfill $K_\alpha\Delta^\alpha \ll L^2$ in the example of the box, where $L$ is the size of the box. For a general confining potential, the turnover time is non-universal and is dependent on all non-zero eigenvalues $\mu_n$ of the Fokker-Planck operator $\mathbb{L}_{\mathrm{FP}}$ \cite{Risken}. Thus, at times $\Delta\gg(1/K_\alpha\mu_1)^{1/\alpha}$ a transition occurs to the $\left< \overline{\delta^2(\Delta)}\right>\sim\Delta^{1-\alpha}$ regime, Eq.~\eqref{msd01}. We stress again that, in contrast to normal diffusion, no saturation is found at long lag times, and $\left<\overline{\delta^2(\Delta)} \right>$ continues to grow for any $\Delta<T$. Only as $\Delta$ approaches to the measurement time $T$, we obtain the convergence $\left<\overline{\delta^2(\Delta)}\right>\to\langle x^2\rangle+\langle x^2(0)\rangle$. \begin{figure} \includegraphics[width=8cm]{figure03.eps} \caption{Single trajectories for the motion in the harmonic binding potential $U(x)=x^2$, for the same parameters as in Fig.~\ref{fig1} (thin lines). The symbols ($\square$) represent the simulations result for $\left<\overline{ \delta^2(\Delta)}\right>$. The scatter between individual trajectories is distinct and resembles qualitatively typical experimental results \cite{Golding,b1,b2,b3,b4,b5}. \label{fig3}}

17 — 1003.3655

\caption{Log-log graphs of magnetization (\opencircle) and energy (\opensquare) autocorrelation time (in $MCS$) versus linear size $L$ for the Niedermayer algorithm with $E_0=-0.9$. The quoted value for $z$ is obtained from the slope of an adjusted straight line for the magnetization autocorrelation time for $L \ge 16$ (see text). The dotted line is just a guide to the eye.}

\caption{Log-log graphs of magnetization (\opencircle) and energy (\opensquare) autocorrelation time (in $MCS$) versus linear size $L$ for the Niedermayer algorithm with $E_0=0.0$. The quoted value for $z$ is obtained from the slope of an adjusted straight line for the magnetization autocorrelation time, for values of $L$ beyond the point where the crossover takes place. The dotted line is just a guide to the eye.}

18 — 1003.3745

\caption{X-ray sources with optical matches within the central $5\arcmin$ of Willman 1. \textdagger\According to\citet{Abazajian:2009a}. \textdaggerdbl\Photometric redshift from\citet{Richards:2009a}. \textasteriskcentered\Photometric classification from\citet{Richards:2009a}}

19 — 1003.3786

\caption{\coloronline Total static structure factor extracted from the simulation. The packing fractions $\varphi=0.45$, $0.5$, $0.55$, and $0.57$ are marked with crosses (red), squares (green), circles (blue) and triangles (magenta), respectively. The black solid line is the Percus-Yevick result for $\varphi = 0.516$. The inset shows a magnification of the peak.}

\caption{\coloronline Partial static structure factors extracted from the simulation by binning the polydisperse system into $M=3$ species, at packing fraction $\varphi = 0.5$. The upper curves show the diagonal terms $S_{11}(q)$ (red crosses), $S_{22}(q)$ (green squares), $S_{33}(q)$ (blue circles). The lower curves show the off-diagonal terms $S_{12}(q)$ (red crosses), $S_{13}(q)$ (green squares) and $S_{23}(q)$ (blue circles).}

\caption{\coloronline Simulation correlators rescaled by the $\alpha$ timescale $\tau \propto t'_\sigma$ for four different wave numbers $q$ as indicated, where $\tau$ was determined by shifting the results for lower $\varphi$ to agree at long times with the $\varphi=0.585$ curve at $qd=7$, and does not depend on $q$. Packing fractions shown are $\varphi = 0.585$ (black plus symbols), $\varphi = 0.58$ (red crosses), $\varphi = 0.57$ (green stars), $\varphi = 0.55$ (blue open squares), $\varphi = 0.53$ (magenta filled squares) and $\varphi = 0.50$ (black circles).}

\caption{\coloronline Fits of stretched-exponential Kohlrausch functions (dashed black lines) to the simulated coherent density correlators (circles) at $\varphi =0.585$. The q-values are from top to bottom $qd= 6.6$ (red), $7.4$ (black), $9.8$ (blue), $12.8$ (green) and $15.6$ (magenta). The fit range was chosen as $t\in[100:10^5]$. Solid black lines show von~Schweidler fits up to second order, Eq.~\eqref{vsfits}, with $t_\sigma =1000$ and $b=0.53$, fitted in the range $t\in[20:336]$.}

\caption{\coloronline MCT results for the critical nonergodicity parameters $f^c_q$ obtained by binning the simulated static structure-factor data into $M=1$ (black dashed), $M=3$ (red solid) and $M=5$ components (blue dash-dotted). The grey dotted curve is the MCT solution with the Percus Yevick static structure factor ($M=1$). Circles show the plateau values of the simulation curves obtained by fitting Kohlrausch functions to the coherent simulation correlators at $\varphi=0.585$ (Fig.~\ref{kohl}). Triangles are obtained by von~Schweidler fits (Eq.~\eqref{vsfits} and Sec.~\ref{beta-process}). % \CHANGE{$M=1$, $3$, $5$ as black dashed, red solid, blue dash-dotted. % PY as grey dotted. Open triangles.} }

\caption{\coloronline MCT results for the tagged-particle critical nonergodicity parameters $f^{s,c}_q$ for the $M=1$ (dashed), $M=3$, (solid) and $M=5$ (dash dotted) polydispersity models. The solutions for $M=3$ and $M=5$ cannot be distinguished on the scale fo the figure. Open circles and triangles show plateau values determined from fitting Kohlrausch and von~Schweidler functions, respectively, as in Fig.~\ref{fqall}. % \CHANGE{symbols, lines and colors corrsponding to Fig.~\ref{fqall}} \label{fq_inc} }

\caption{\coloronline $\alpha$-relaxation times $\tau_q$ obtained by stretched-exponential fits to the MCT master curves for the $M=1$ (black dashed), $M=3$ (red solid), and $M=5$ (blue dash-dotted) polydispersity approximations. All values have been rescaled by their value at $q_pd=7.3$, $\tau^*=\tau_q/\tau_{q_p}$. The upper panel shows data corresponding to the collective density correlation functions, the lower panel those corresponding to the tagged-particle analog. Circles correspond to the $\tau_q$ extracted from the simulation data at $\varphi=0.585$, as in Fig.~\ref{fqall}. % \CHANGE{symbols, line styles, colors as in Fig.~\ref{fqall}} }

\caption{\coloronline Kohlrausch stretching parameters $\beta_q$ as functions of wave number $q$, determined from fits to the MCT $\alpha$-master curve for the $M=1$ (dashed), $M=3$ (solid), and $M=5$ (dash-dotted) polydispersity moment approximations. $\beta_q$ determined from fits to the simulation data (as in Fig.~\ref{fqall}), are shown as open circles. The upper (lower) panel shows results from analyzing the collective (tagged-particle) quantities. % \CHANGE{colors, line styles, symbols as in Fig.~\ref{fqall} + error bars} }

\caption{\coloronline $\beta$ analysis of the simulation data at $\varphi = 0.585$. Functions $X(q,t)$ calculated from Eq.~\eqref{xeq} by fixing $t'=10.0085$ and $t''=52.5810$ are shown for the collective (tagged-particle) correlators in the upper (lower) panel. Different wave numbers $q$ were chosen as labeled. The MCT factorization theorem is validated by observing data collapse for different $q$ in a time window spanning $[t',t'']$, and by the $\beta$-master curve shown as a dashed line. \label{xfunctions} }

\caption{\coloronline Critical amplitudes $h_q$ calculated within MCT for the $M=1$ (dashed), $M=3$ (solid), and $M=5$ (dashed-dotted) moment approximations to the simulated polydispersity distribution. For comparison, the PY-based theoretical result is also shown (cyan line). Open circles and triangles mark the amplitudes determined from the $\varphi=0.585$ simulation data via the function $Y(q)$, Eq.~\eqref{Yfunc}, and via the von~Schweidler fits discussed in conjunction with Fig.~\ref{fqall}, respectively. The results for $Y(q)$ have been scaled by a factor $0.61$ to account for the unknown normalization in this procedure. %\CHANGE{symbols, colors, lines as in Fig.~\ref{fqall}; ordinate labeled $h(q)$, not $h^c(q)$} }

\caption{\coloronline Critical amplitudes $h_q^s$ calculated within MCT for $M=1$ (dashed), $M=3$ (solid), and $M=5$ component (dashed-dotted) approximations to the polydispersity distribution, and by using the Percus-Yevick static structure factor (cyan). Open circles are the corresponding amplitudes determined from $Y(q)$, Eq.~\eqref{Yfunc}, scaled by a factor $0.63$. Triangles show results from von~Schweidler fits as in Fig.~\ref{fqall}. %\CHANGE{all styles as in Fig.~\ref{fqall}} }

\caption{\coloronline Packing fractions $\varphi_\text{MCT}$ chosen for the five-component (triangles), three-component (circles), and one-component (squares) MCT fits to the simulation correlators, as functions of the simulated packing fraction $\varphi$. The dashed lines are linear regression fits, $\varphi_\text{MCT}=a\varphi+b$, with parameters $a=1.0001$ ($0.9506$, $0.9083$) and $b=-0.0285$ ($0.0288$, $0.0288$) for $M=1$ ($M=3$, $M=5$). Crosses mark $\varphi_\text{MCT}$ from a fit to the mean-squared displacement only ($M=3$). Horizontal dotted lines show the MCT predictions for the critical point for each of the three models; their intersections with the linear regression curve marks the estimated $\varphi^c$ for the simulation, as noted by arrows on the abscissa.}

\caption{\coloronline MCT fits of the simulated collective density correlation functions (shown as circles) using the simulated static structure factors binned into $M=1$ (dashed), $M=3$ (solid), and $M=5$ (dash-dotted lines) components to approximate the simulated polydispersity distribution. Packing fractions in the simulation are $\varphi=0.5$, $0.53$, $0.55$, $0.57$, $0.58$, and $0.585$. The curves have been fitted by adjusting only $\varphi_\text{MCT}$ as described in conjunction with Fig.~\ref{critpointMD}; we get $\varphi_{\text{MCT},M=1}=0.473$, $0.502$, $0.52$, $0.54$, $0.554$, $0.558$; $\varphi_{\text{MCT},M=3}=0.449$, $0.472$, $0.493$, $0.5122$, $0.5234$, $0.5289$; and $\varphi_{\text{MCT},M=5}=0.45$, $0.47$, $0.49$, $0.509$, $0.5207$, $0.5259$. %\CHANGE{all styles as in Fig.~\ref{fqall}} }

\caption{\coloronline MCT description of the collective density correlation functions from the simulation at $qd = 12.8$ (corresponding to the second peak of the averaged static structure factor). Packing fraction and symbols as in Fig.~\ref{fitatpeak}. % \CHANGE{\textellipsis\ as in Fig.~\ref{fqall}} }

\caption{\coloronline MCT description of collective density correlation functions from the simulation at $qd=9.5$ (corresponding to the first dip after the main peak of the averaged static structure factor). Packing fraction and symbols as in Fig.~\ref{fitatpeak}. %\CHANGE{\textellipsis as in Fig.~\ref{fqall}} Arrows on the right mark the plateau values obtained from Kohlrausch fits discussed in Fig.~\ref{fqall}. The two curves starting at the plateau values are the solutions to the corresponding $\alpha$-master equation.}

\caption{\coloronline MCT description of collective density correlation functions from the simulation at $qd=3.0$, with symbols and packing fractions as in Fig.~\ref{fitatpeak}. The inset shows only the highest packing fraction, $\varphi=0.585$, and its corresponding $M=1$ correlator, in order to highlight the different relaxation times of the different MCT approaches. %\CHANGE{inset: only one, not three densities, inset slightly smaller; %same color code in inset as in main figure, adapt to Fig.~\ref{fqall}} }

\caption{\coloronline MCT description of the tagged-particle density correlation functions determined from the simulation at $qd=7.3$; packing fractions and symbols as in Fig.~\ref{fitatpeak}.%\CHANGE{\textellipsis\ as in Fig.~\ref{fqall}} }

\caption{\coloronline MCT description of the tagged-particle density correlation functions as in Fig.~\ref{incoherent}, but for $qd=9.5$. %\CHANGE{\textellipsis as in Fig.~\ref{fqall}} }

\caption{\coloronline Mean squared displacement from the simulation (open circles) compared with their MCT description based on the fit performed in Fig.~\ref{fitatpeak}; different line styles correspond to $M=1$, $M=3$, and $M=5$ models of the polydispersity distribution, cf.\Fig.~\ref{fitatpeak}. Arrows indicate the times for which the van~Hove functions are shown in Fig.~\ref{vHove_log_lowpack} (dashed) and Fig.~\ref{vHove_log} (solid arrows). %\CHANGE{\textellipsis\ as in Fig.~\ref{fqall}} }

\caption{\coloronline $P(\log \delta r) = \ln(10) 4 \pi \delta r^3 G_s(r,t)$ for the times marked by the dashed arrows in Fig.~\ref{msd}. Open circles are the simulation results for $\varphi=0.5$, while lines indicate the corresponding MCT result for the $M=3$ multi-component model at $\varphi_\text{MCT}=0.449$: solid black lines denote the full distribution, while red-dashed (cyan-short-dashed, blue-dashed-dotted) mark the distributions involving only small (medium, large) particles. %\CHANGE{colors, lines!} }

\caption{\coloronline $P(\log \delta r)$ as in Fig.~\ref{vHove_log_lowpack}, for the times marked by solid arrows in Fig.~\ref{msd} and for packing fraction $\varphi=0.585$ (fitted by $\varphi_{\text{MCT},M=3}=0.5289$). The inset shows the average MCT distributions with an additional result at $t=1.01 \; 10^{5}$ where long-time diffusion has already set in. The dotted grey line is a fit with a Gaussian distribution, $(4\pi Dt)^{-3/2}\exp[-\delta r^2/(4Dt)]$, where $D=8.35\times10^{-6}$ is taken from Fig.~\ref{msd}.}

\caption{\coloronline $\alpha$-relaxation time scale $\tau$ determined from the master curve of Fig.~\ref{rescaled-cor} (red circles), and inverse long-time self-diffusion coefficients $1/D$ extracted from the simulated mean-squared displacements (green diamonds). The red-solid line shows the power law $\tau \sim|\epsilon|^{-\gamma}$ with $\gamma=2.445$ determined from the $M=3$ MCT calculations. %The fitting factor for $\tau_s = A |\epsilon|^{-\gamma}$ is $A=0.025$. % A2 = 0.0272116524097195 Inset: product $D\tau$ evaluated from the simulation data (black up-triangles) and from MCT (red down-triangles) scaled by a factor to make them comparable. %\com{mf, ist dies noetig? wie falsch ist der MCT wert?} % 50 !!!! % \CHANGE{$D\tau$ from MCT scaled by a factor so that we can compare; % added same power law with smaller prefactor so % that it fits the last two or three red circles} }

\caption{\coloronline Upper panel: Inverse self Diffusion coefficient $1/D$ (green diamonds) and $\alpha$-relaxation time scale $\tau$ determined from the master curve of Fig.~\ref{rescaled-cor} (red circles) with double logarithmic axes. Lower panel: ${1/D}^{-1/\gamma}$ (green diamonds) and $\tau^{-1/\gamma}$ (red circles) from the upper panel. The errobars are estimated from the different $\gamma$-values obtained from the MCT calculations for $M=1,3,5$ components. In both panels the solid red line shows the corresponding power law from Fig.~\ref{powerlaw}. % \CHANGE{rectification plot... Diffusion is less linear! Dashed black line: powerlaw with $\gamma = 2.445$ as in fig \ref{powerlaw}, errorbars from variation of $\gamma_{1cp}...\gamma_{5cp}$} }

\caption{\coloronline Relaxation times $\tau^s_q$ of the tagged-particle density correlation functions at packing fraction $\varphi=0.585$ (upper panel; $\varphi_\text{MCT}=0.5289$) and $\varphi=0.5$ (lower panel; $\varphi_\text{MCT}=0.449$), plotted as $q^2\tau^s_q$. Filled symbols (full lines): $\tau^s_q$ extracted from Kohlrausch fits to the simulation (theory) data, taking into account the plateau values $f^s_q$. Open symbols (dotted lines): determined via $\phi^s(q,\tau^s_q)=1/e$. }

20 — 1004.0153

\caption{(Color online) (a) Schematic of QD samples and interferometer. (b) Area-normalized emission lines for QD1 (\textcolor{blue}{$\circ$}) and QD2 (\textcolor{red}{$\blacktriangle$}). (c) Time-dependent fluorescence from QD1. (d) Time-dependent fluorescence from QD2.}

\caption{(Color online) (a) Correlation of the interference for orthogonal polarizations with simulated curve. (b) Correlation of the interference for parallel polarizations with simulated curve. (c) Close-up of $\tau=0$ peak for orthogonal (\textcolor{red}{$\vartriangle$}) and parallel (\textcolor{blue}{$\bullet$}) polarizations. The solid and dashed curves are simulations including the detectors' time response, while the dash-dotted curve is the expected curve for infinitely fast detectors.}

21 — 1004.0965

\caption{\label{Fig2}Low temperature size calibration of self-assembled atomic clusters in the model described in the text. The number of clusters of different sizes at different temperatures: \opendiamond---$T=|v_1|$, +---$T=0.1|v_1|$ and \opencircle---$T=0.01|v_1|$. The curves are guides to the eye except the monotonous dashed curve which describes random coverage.}

22 — 1004.3300

\caption{(Color) Magnetization vs.\external field (DMRG data) for different$n$ as shown in the inset, for $\alpha=J_2/|J_1|=1/3$, $\beta_2=0.05$, $\beta_1=0$, and $L=24$ sites in each chain. Lower inset: The 3D arrangement of chains used for the DMRG study reported here. Upper inset: Magnetization curve inside the 3-MBS region for $\beta_2=0.005$ (blue curve, \textcolor{blue}{$\times$}) compared with that for the 1-magnon C-phase for $\beta_2=0.05$ (red curve, \textcolor{red}{$\circ$}). Note the three times larger step for $\beta_2=0.005$. }

\caption{(Color) Upper: Approximate relative in-chain contribution to the saturation field measured by its deviation from Eq.\(1) for different $\alpha$ values vs.\$\beta_2$ in units of its critical value $\beta^{\rm cr,+}_2$ (see Figs.\3, S1, and text). The yellow stripe highlights the region of predominant IC. The full line parts of a curve (below the symbol$\diamond$) can be determined within the SWT, whereas the branches given by symbols are obtained by the DMRG and the HCBA methods. Lower: in-chain and IC spin-spin correlation functions vs.\applied field, given by the filled symbols{\color[named]{Orange}$\blacksquare$} and {\Large \color[named]{Green}${ \bullet}$}, respectively, as obtained by DMRG. }

\caption{(Color) Saturation field for different in-chain frustration rates $\alpha=0.5$, $0.75$, and $1.0$ vs.\IC$\beta_2/\alpha=J_{\rm ic}/J_2$. The results of the exact hard-core boson approach are shown by symbols; the red solid line shows the one-magnon result, which becomes exact for $\beta_2 > \beta^{\rm cr,-}_2=J^{\rm cr,-}_{\rm ic}/|J_1|$ as explained in the main text (see also Figs.\3-5, therein). The green dashed line shows the value, which is reached in the commensurate C-phase, where the dependence on$J_1$, $J_2$ vanishes identically; the magenta thin line shows an approximation to the exact solution in the 2-magnon phase. %Schematic high-field phase diagram of %the model (\ref{Hviab}), the labels C, INC1, and INC2 denote the region of the usual %one-magnon dipolar phase, the "OP"-quadrupolar and the "OP"-octupolar phases, %respectively. }

23 — 1004.3501

\caption{\label{fig:PELcomparison} % The speedup and memory cost of using a MinMax PEL of various sizes relative to storing only a single event per particle. The dashed lines are the values for using a STL priority\_queue. The results are for a monodisperse system of$N=10^5$ elastic hard spheres simulated on an Intel\textregistered\Core\texttrademark i5 750 with 8~MB L3 cache and 8~GB of RAM.}

\caption{\label{fig:CoilSnapshot} % A snapshot of $5\times10^5$ sheared inelastic hard spheres, generated live from a simulation, displaying a characteristic clustering behavior of granular systems. On a NVIDIA\textregistered\Geforce GTX 260\texttrademark, the image is updated at 15 frames per second. }

\caption{\label{fig:scaling} % The number of collision events processed per second using linear neighbor lists $N_{\rm coll}^{(\rm linear)}$ as a function of (a) the number of particles $N$ and (b) the number density $\rho$. The lines indicating the cache memory boundary is approximate as different densities incur slightly different memory requirements. The results are for a monodisperse system of elastic hard spheres simulated on an Intel\textregistered Core\texttrademark i5 750 with 8~MB L3 cache and 8~GB of RAM.}

24 — 1004.3851

\caption{(color online) Plot of the reduced mobility $F_S^*$ as a function of the reduced interaction $\Gamma$ for colloids with magnetic dipole interactions. \textcolor{blue}{$\bullet$}~: Numerical simulations \cite{Herrera08}, \textcolor{red}{$\blacksquare$}~: experimental data \cite{Wei00}, solid line Kollmann's formula, Eqn.~(\ref{eq:mobreduced}) with (\ref{eq:gammapara}) (no free parameter in this calculation).}

\caption{(color online) Plot of the reduced mobility $F_S^*$ as a function of the brownian particles packing fraction $\varphi$, for colloids with screened electrostatic interactions (DLVO). \textcolor{blue}{$\bullet$}~: Numerical simulations \cite{Herrera07}, solid line Kollmann's formula, Eqn.~\refpar[eq:mobreduced] with \refpar[eq:gammaDLVO] (no free parameter in this calculation).}

25 — 1004.3957

\caption{(Color online) Incommensurate signals in \csreda. (a) and (b): diagonal scans across \Qab, the thin dotted line is the scattering angle dependent background. (c) and (d): transverse scans (background subtracted) across (1,0,0), crossing the positions \Qice\and\Qicz\at two temperatures. Scans were taken with an energy transfer of 4~meV on the 1T spectrometer.\label{icfig1}}

\caption{(Color online) Magnetic scattering in \csreda\taken around the FM wave vector (0,0,1.6) on the 4F spectrometer. (a): Constant energy scans at different temperatures. Shifted by 2 units each. (b): Signal at (0,0,1.6) as function of temperature. (c): Signal as function of energy at different temperatures (shifted by 100$\mu_B^2/eV$ each). Lines are fits to a single relaxor function (see text). \label{fmfig1}}

\caption{(Color online) Analysis of FM scattering in \csreda. (a): Characteristic energy, as obtained from the fits in Fig.\\ref{fmfig1}. (b): Value of the real part of the macroscopic susceptibility, calibrated to absolute units, obtained from the same fits. (Inset: inverse of the same values.) \label{fmfig2}}

\caption{(Color online) Magnetic scattering in \csreda\(T=0.05~K, shifted by 20 counts each). The model used to fit the data in (a) consists of a ferromagnetic contribution and two incommensurate AFM-like excitations at \Qice\and\Qicz. These components are given by equations (\ref{eqrelaxor})-(\ref{eqanti}) and are displayed separately in part (b) and (c). The FM one has to be modified to account for the measured finite energy at the zone center. Data taken on the IN14 spectrometer.\label{fitfig1}}

\caption{(Color online) Ferromagnetic fluctuations in \csreda: different data sets taken at constant Temperature or energy transfer (same as in Figure \ref{fmfig1}). The quantity $\omega\chi''$ only depends on $\omega/(T+\Theta)$. The line is the function $f$ for the parameter values discussed in the text. \label{scalfig}}

\caption{(Color online) Magnetic scattering in \csreda \as a function of temperature and energy (background subtracted, shifted by 300 counts each). All constant energy scans along$Q$=(1,K,0) were performed on the thermal triple-axis spectrometer 1T. The signal changes its shape qualitatively due to the interplay of ferromagnetic and incommensurate contributions. At high energy and low temperature the incommensurate signals dominate, whereas at low energy and high temperature the centered peak in these scans arises from the ferromagnetic component.}

\caption{(Color online) Effect of a magnetic field on the ferromagnetic fluctuations in \csreda\(a) and \csrea\(b). Temperature is 1.5~K and background has been subtracted in both cases (but the scales are not normalized to each other). Lines correspond to a single relaxor (B=0, with the characteristic energies discussed in the text) and to Lorentzian functions (B=10~T), respectively. Data were taken on the IN12 and Panda spectrometers.\label{magfig1}}

26 — 1004.3994

\caption{\textcolor{red}{Small step} or \textcolor{blue}{giant leap}? For this potential, the answer is small step.}

\caption{What goes up must come down. In interpolating between the two vacua, the field in the wall does not follow a \textcolor{red}{straight line} in field space, instead it \textcolor{blue}{loops} up to large $y$ and then back down to $y=1$.}

27 — 1004.4368

\caption{\small Event distribution in the invariant-mass interval 4.4--6.0~GeV, for the exclusive production of $D^{\ast}\bar{D}^{\ast}$ in initial-state-radiation events, from $e^{+}e^{-}$ annihilations at a center-of-mass energy near 10.58~GeV, as published by the BaBar Collaboration \cite{PRD79p092001}. With ({\color{blue} X}) on the horizontal axis we indicate the harmonic-oscillator vector levels for the parameters $m_{c}=1.562$ GeV and $\omega =0.19$ GeV \cite{PRD27p1527}. Meson loops, first introduced in Ref.~\cite{PRD21p772}, shift the central masses of the $S$ and $D$ charmonium resonances to the positions indicated in the figure. One may observe that all enhancements, with the exception of the one above the $\Lambda_{c}^{+}\Lambda_{c}^{-}$ threshold, correspond to the predicted central mass positions. In the inset we show independent Breit-Wigner fits to each of those resonances that had not yet been firmly determined in previous work.}

28 — 1004.4535

\caption{Mean velocity profile $\Up$ plotted against $\yp$. {\textcolor{red}{$\bullet$}} Experimental data measured with PIV. {\textcolor{blue}{\boldmath{$\relbar\!\relbar\!\relbar$}}} best fit. {\textcolor{black}{\boldmath{$\cdots\,\cdots$}}} Analytical law \unboldmath{$\Up=\log(\yp)/0.41+5$}.}

\caption{Spanwise distribution of the mean velocity $U(x=4.\lz,Y=k/2,z/\lz)/\uinf$. \boldmath{$\circ$} PIV measured velocity distribution. {\textcolor{red}{\boldmath{$\relbar\!\relbar\!\relbar$}}} Best fit to the experimental data with a sinus function.}

29 — 1004.5309

\caption{(color online) Plot of the penetration field $H_{p}^{45}$ of a Bi$_{2}$Sr$_{2}$CaCu$_{2}$O$_{8}$ crystal with the edges cut at $\alpha = 45^{\circ}$ with respect to the main crystal axes (\color{red} $\diamondsuit$ \color{black}), compared to the interpolated values $H_{p}^{0}$ for a crystal with the same $d/w = 0.12$, but $\alpha = 0$ (\color{magenta} \fbox{\hspace{0.3mm}}{} \color{black}). Also shown is $H_{p}^{0\infty}(T)$ for a crystal with $d/w \rightarrow \infty$, $\alpha = 0$, extracted from the fits in Fig.~\protect\ref{fig:tanh} (\color{green} \fbox{\hspace{0.mm}}{} \color{black}); the drawn line is a fit to the theoretical temperature dependence of $H_{c1}$ for a $d$-wave superconductor,\protect\cite{Radtke96} with $H_{c1}(0)= 180$ Oe and $T_{c} = 88.7$ K. The inset shows the experimental (lefthand axis) and the theoretically expected enhancement (righthand axis) of the field of first flux penetration for $\alpha = 45^{\circ}$ with respect to $\alpha = 0$. }

30 — 1005.1935

\caption[Effect of torque on $\alpha$ of LiHoF$_{4}$.]{\label{fig:torque}Measurement of $\alpha$ for LiHoF$_{4}$ at $\pm2.5~T$ in the maximum (solid lines) and minimum (dashed lines) demagnetization orientations. Crosses show the calculated critical temperatures. The \textcolor{red}{red} (\textcolor{blue}{blue}) dotted line indicates the critical temperature determined from the \textcolor{red}{maximum} (\textcolor{blue}{minimum}) demagnetization orientation, identical within error.}

31 — 1005.3951

\caption{Heading (northing) of an acoustic storey (top-most on IL07) from compass data (labelled {\color{blue}$\times$}) and from the positioning of the acoustic sensors (labelled {\tiny{\color{red}$+$}}). The oscillating pattern in these five days is due to well-known Coriolis-induced current variations.}

32 — 1005.4202

\caption{Saturated densities. Simulation results:\textcolor{red}{\large $\circ$}~this work, \textcolor{green}{\small $\square$}~EPM2 \cite{Harris1995}, \textcolor{blue}{\small $\vartriangle$}~Vrabec et al. \cite{Vrabec2001}, \textcolor{blue}{\small $\square$}~Zhang and Duan \cite{Zhang2005,Merker2008}, \textcolor{green}{\small $\vartriangle$}~BBV \cite{Bock2000,Bratschi2007}, {\small $\square$}~M\"{o}ller and Fischer \cite{fischer1994}, {---}~EOS \cite{wagner}, {\small $\bigstar$}~experimental critical point \cite{wagner}. The inset is a magnified view of the critical point.}

\caption{Saturated densities of cyclohexanol: \textcolor{red}{\large $\circ$}~model after first step of optimization, \textcolor{green}{\large$\circ$}~model after reduced unit method, {\large$\circ$}~final model, {---}~DIPPR correlation \cite{DIPPR2005}, $\star$~experimental critical point \cite{DIPPR2005}. Inset: Magnified view on the critical point.}

33 — 1005.4649

\caption{\label{fig:ellipse}(a) {\color{red}Fast (solid line)} and {\color{green}slow (dashed line)} characteristic modes. The fast mode has a major axis perpendicular to $\vec{B_{\perp}}$ and right-handedness with respect to $\vec{B_{\parallel}}$. `$\perp$' and `$\parallel$' are defined with respect to the propagation direction. (b) Polarization properties are characterized by $\chi$ (\textit{elliptization angle}, where right/left handedness are represented by $+/-$ sign) and $\psi$ (\textit{polarization direction angle}); their ranges are also shown.}

\caption{\label{fig:profiles}(a) Plasma density profile of base case for modeling. (shot \#$124764$, $0.325\ sec$). (b) Toroidal ({\color{blue}solid line}) and vertical ({\color{green}dashed line}) magnetic field along major radius in the mid-plane (they vary little with height near the mid-plane). (c) Horizontal (i.e.\radial) magnetic field along major radius$0.1\ m$ above ({\color{blue} blue dashed line}), below ({\color{green} green dashed line}) and in the mid-plane ({\color{red} red solid line}).}

\caption{\label{fig:chi}Elliptization angle ($\chi$) evolution along chord in mid-plane for waves with horizontal linear polarization at launch (i.e.\in toroidal direction). Vertical solid line indicates plasma center (i.e.\peak density). The mirror is mounted on the center stack. ($f = 288\ GHz, \lambda = 1.04\ mm$)}

\caption{\label{fig:chi_0_45}Elliptization angle ($\chi$) evolution for horizontal ({\color{red} solid line}) and $45^{\circ}$ ({\color{green} dashed line}) linear polarization at launch along chord in mid-plane. Zero elliptization (i.e.\linearly polarized) is highlighted on the grid. ($f = 288\ GHz, \lambda = 1.04\ mm$)}

\caption{\label{fig:real_CM}Elliptization angle ($\chi$) evolution of wave with horizontal linear polarization at launch along a chord $0.1\ m$ above mid-plane with ({\color{red} solid line}) and without({\color{green} dashed line}) Faraday rotation. Faraday rotation is eliminated by setting $\vec{B_{\parallel}} = 0$ along chord). ($f = 288\ GHz, \lambda = 1.04\ mm$)}

34 — 1005.4740

\caption{ \label{fig_5} (Color online) Scaling of the shift of the finite-size transition points, calculated from the mean value ({\color{red} $+$} for 'b' and {\color{blue}$\rlap{$\times$}{+}$} for 'f'), as well as from the median ({\color{green}$\times$} for 'b' and {\color{magenta}$\boxdot$} for 'f') of the distribution, as a function of $L$ in log-log scale for both type of disorder. Estimates for $\theta_c$ are taken from Eq.(\ref{true_cp}) and the straight lines indicating the asymptotic behavior have the same slope: $-1/\nu_s \approx -0.8$. The effective critical exponents calculated from Eq.(\ref{eff_shift}) are shown in the upper inset (for the mean) and in the lower inset (for the median).}

35 — 1005.4745

\caption{\label{fig4} Local spin polarization at (a)-(c) three representative energies. Raw- and {\color{blue}avg} (over multiple images and according to symmetry\\cite{EPAPS}) experimental data are compared with DFT simulations incl.\{\color{red}vdW} and incl.\{\color[rgb]{1,0,1}{S}}pin-{\color[rgb]{1,0,1}{O}}rbit-{\color[rgb]{1,0,1}{C}}oupling. Line profiles follow high-symmetry directions as indicated in the sphere model inset in (a). Simulated data including SOC is only given in (b) as the SOC corrections for (a) and (c) are negligible. Insets: $22~\text{\AA}\times 22~\text{\AA}$.}

36 — 1006.0013

\caption{(color online) The function $g(\omega)$, given by the RHS of Eq.~(\ref{brewster}), as a function of frequency for TE (full curve) and TM (dashed curve) waves. For a given angle of incidence, $\theta$, the Brewster anomalies occur at frequencies satisfying $\cos^2\theta=g(\omega)$. As an example, the horizontal dotted line corresponds to $\cos^2\pi/6$, whose intersections (indicated by arrows) with $g(\omega)$ yield the frequencies for Brewster anomalies for $\theta = \pi/6$. The media parameters are the same as in Fig.~\ref{fig1}. \color{red} \color{black} }

\caption{(color online) Localization length $\xi$ (in units of the system size) for TE (a) and TM (b) waves as a function of the vacuum wavelength $\lambda$, obtained from our numerical simulations, for different angles of incidence (solid lines, from top to bottom): $\theta = 0$, $\pi/100$, $\pi/12$, $\pi/6$, and $\pi/3$. Vertical arrows locate the Brewster modes, and the dashed line corresponds to the asymptotic behavior $\xi \propto \lambda^{6}$ predicted in Ref.~\cite{asatryan}. The media parameters are the same as in Fig.~\ref{fig1}, with $\Delta=0.5$ mm. \color{red} \color{black} }

37 — 1006.0635

\caption{Typical trajectories of the electrons (\full) and positrons (\dashed) under planar channeling (left) and above-barrier motion (right). Pluses mark the positions of atomic strings (perpendicular to the plane of the figure) forming the atomic planes of the crystal.}

\caption{Incoherent bremsstrahlung efficiency (in ratio to the Bethe-Heitler efficiency in amorphous medium) from 1 GeV electrons (\full) and positrons (\dashed) vs incidence angle $\theta$ to $(0\bar 11)$ plane of Si crystal, as a result of simulation.}

38 — 1006.0933

\caption{Magnetic field dependence of the differential conductance peak at bias +1.4~mV. Filled circles (\textcolor[rgb]{0.00,0.00,0.50}{\Circsteel}) correspond to the 90$^\circ$ angle between \textbf{$\vec{B}$} and the plane of the heterointerface. Open circles (\textcolor[rgb]{0.00,0.40,0.29}{\Circpipe}) are the data obtained at an angle of 0$^\circ$. Solid curve and dotted curve depict the best fit achieved.}

39 — 1006.1906

\caption{% \label{f:analyt}% (color) MOKE signals $|\Phi|$ originating from the different types of spin waves as a function of the Ni thickness, assuming the spin-wave ellipticity $\epsilon=2$. (\textcolor{red}{$\blacktriangledown$}, \textcolor{green}{$\blacktriangle$}) denotes for the DE modes bounded to upper and lower FM interface, respectively, with $\kswz=10^{7}$\,m$^{-1}$. (\textcolor{magenta}{$\star$}) denotes for the FMR mode (i.e.\as the DE modes with$\kswz=0$). (\textcolor{blue}{$\blacksquare$}) denotes for the PSSW modes. Incidence angle is 25$^\circ$, light wavelength is 810\,nm. For values of optical and magneto-optical parameters, see text. Symbols are the optically exact$4\times4$ matrix formalism, solid lines the analytical expressions (Eqs.~(\ref{eq:MOKEde}, \ref{eq:MOKEdebot}, \ref{eq:MOKEpssw})). }

40 — 1006.2125

\caption{(color online) (Left) Fraction of infected nodes $\left<I(t)/N\right>$ as a function of time for the original event sequence (\textopenbullet) and null models: equal-weight link-sequence shuffled DCWB ($\lozenge$), link-sequence shuffled DCB ($\vartriangle$), time-shuffled DCW ($\square$) and configuration model D ($\triangledown$). Inset: $\left<I(t)/N\right>$ for the early stages, illustrating differences in the times to reach $\left<I(t)/N\right>=20\%$. (Right) Distribution of full prevalence times $P(t_f)$ due to randomness in initial conditions.}

41 — 1006.2954

\caption{$g$-dependence of particle populations in the symmetric stationary states. $\Box$($+$) represents $N_{a}\left(N_{b}\right)$.} \label{fig:fig_2} \end{figure} \begin{figure}[htbp] \includegraphics[width=8.3cm]{Fig_3}% Here is how to import EPS art \caption{$g$-dependence of excitation spectra of the symmetric stationary states.} \label{fig:fig_3} \end{figure} \begin{figure}[htbp] \includegraphics[width=8.3cm]{Fig_4}% Here is how to import EPS art \caption{$g$-dependence of imaginary part and real part of $\hbar \omega_{-}$ at symmetric stationary states.} \label{fig:fig_4} \end{figure} In Fig. \ref{fig:fig_2}, we investigate the $g$-dependence of particle populations at the symmetric stationary state. From this figure we conclude that the particle populations are large enough for applying the mean-field approximation in a wide range of atom-molecule internal tunneling $g$. We also find that the atomic populations in the ground states grows, by increasing the atom-molecule internal tunneling strength. In the symmetric stationary state, where $2 \theta_{aL(aR)} - \theta_{bL(bR)} = 0$, $N_{a} = N_{aL(aR)}$ and $N_{b} = N_{bL(bR)}$, the internal tunneling term in the Hamiltonian (\ref{eq:classical_Hamiltonian}) is reduced to be $- 4 g N_{a} \sqrt{ N_{b} }$. Because the order of $N_{a}$ is larger than that of $N_{b}$, the symmetric stationary state tends to lower the total energy by increasing $N_{a}$ rather than $N_{b}$ in the large $g$ region. \begin{figure}[htbp] \includegraphics[width=8.3cm]{Fig_5}% Here is how to import EPS art \caption{$g$-dependence of atomic particle populations at symmetric(+ points) and asymmetric stationary states($\bigtriangleup$ and $\Box$ points).} \label{fig:fig_5} \end{figure} \begin{figure}[htbp] \includegraphics[width=8.3cm]{Fig_6}% Here is how to import EPS art \caption{$g$-dependence of molecular particle populations at symmetric(+ points) and asymmetric stationary states($\bigtriangleup$ and $\Box$ points).} \label{fig:fig_6} \end{figure} Next, we investigate the dynamical stability of the symmetric stationary state by looking at the excitation frequencies. Fig. \ref{fig:fig_3} is the excitation spectra. $\omega_{+}$ and $\omega_{AM}$ are always real and close to each other in the large $g$ region. On the other hand, as shown in Fig. \ref{fig:fig_3}, $\omega_{-}$ goes to zero at the finite atom-molecule internal tunneling $g$. As shown in Fig. \ref{fig:fig_4}, the imaginary part of $\omega_{-}$ emerges at the same value of $g$, while the other modes are still dynamically stable. This fact indicates the occurrence of symmetry-breaking phase transition. \begin{figure} %\begin{minipage}{9pc} \includegraphics[width=8.3cm]{Fig_7}% Here is how to import EPS art \caption{$g$-dependence of total energy at symmetric and asymmetric stationary states. The inset is energy difference between symmetric and asymmetric stationary states. $E_{s} \left(E_{a} \right)$ represents the total energy of symmetric(asymmetric) stationary state.}

42 — 1006.3701

\caption{Quality analysis of RAE results for physics. (a) Quality scores normalised to the overall discipline average given by (\ref{intensive}) for alphabetically listed teams. (b) The same data plotted against the sizes $N$ of research teams where the solid curve is the fit coming from the model (\ref{extensive}) and the dashed curves are the corresponding 95\% confidence intervals. The dotted line indicates how the left trend line would continue if the breakpoint (upper critical point) were absent. (c) Quality renormalised to the expectation values $\langle{ s }\rangle$ coming from the model (\ref{extensive}). The various symbols indicate members of the Russell Group~({\color{red}{+}}), the 1994 Group~({\color{green}{$\times$}}), the Million+ Group~({\color{Gray}{$\blacktriangle$}}), the University Alliance~({\color{blue}{$\Diamond$}}), %Guild HE~({\color{green}{$\blacktriangledown$}}), and unaffiliated universities~({\color{magenta}{$\Box$}}). }

\caption{Quality analysis for geography, Earth and environmental sciences analogous to Fig.1. The standard deviation reduces from $12.8$ in panel (a) to $7.5$ in panel (c). The various symbols (in colour online) indicate members of the Russell Group~({\color{red}{+}}), the 1994 Group~({\color{green}{$\times$}}), the Million+ Group~({\color{Gray}{$\blacktriangle$}}), the University Alliance~({\color{blue}{$\Diamond$}}), Guild HE~({\color{green}{$\blacktriangledown$}}), and unaffiliated universities~({\color{magenta}{$\Box$}}).}

43 — 1006.3985

\caption{Selection fractions of metal-poor candidates selected with the HES selection criteria as described in Paper~IV. The two panels correspond to selection efficiencies using KP and \deredBV ~(left) or \deredJK ~(right). Different lines refer to different red cutoffs as shown in the legend; the solid lines refer to the total selection fractions.}

44 — 1006.4087

\caption{(Color online). Optical depth as a function of density for 60$S_{1/2}$ (\textcolor{blue}{$\blacktriangleright$}, \textcolor{blue}{$\bullet$}) and 54$S_{1/2}$ (\textcolor{red}{$\vartriangleright$}, \textcolor{red}{$\circ$}) for weak and strong $\Omega_\mathrm{p}$ respectively, scaled by probe only optical depth to remove trivial linear scaling. Strong probe data reveals second order density scaling consistent with a cooperative optical non-linearity. Comparison of the gradients for $60S_{1/2}$ and $54S_{1/2}$ gives a ratio of $2.6\pm0.7$.}

45 — 1007.0015

\caption{Detector sensitivities (90\% C.L.) for {\melissa} and VT-1 for a Teflon{\textregistered} ring.}

46 — 1007.0161

\caption{\label{fig:energy} (Color online) Energy per particle of the commensurate system ({\color{red}$\lozenge$}) and the incommensurate system with one of the 180 sites being vacant ({\color{blue} $\circ$}). Error bars are smaller than the symbols and are shown within them. Solid lines in the main plot area are third-degree polynomial fits as described in the text. Lower inset separately shows statistical errors with solid symbols. Top inset shows pressures for both systems (pressure of the incommensurate solid is higher). Dashed line marks melting pressure, below which both systems are metastable but remain crystalline.}

47 — 1007.1155

\caption{Typical trajectories of the electrons (\full) and positrons (\dashed) under planar channeling (left) and above-barrier motion (right). Pluses mark the positions of atomic strings (perpendicular to the plane of the figure) forming the atomic planes of the crystal.}

\caption{Incoherent bremsstrahlung efficiency (in ratio to the Bethe-Heitler efficiency in amorphous medium) from 1 GeV electrons (\full) and positrons (\dashed) vs incidence angle $\theta$ to $(0\bar 11)$ plane of Si crystal, as a result of simulation.}

\caption{Coherent (\chain) and incoherent (\full) contributions to the relative efficiency of radiation (\dashed) according to simulation, in comparison to the experimental results \cite{Backe} (thick curve).}

48 — 1007.2210

\caption{(Color online) Calculated spectra (ABINIT LDA, NOSO) of $\epsilon_{1,2}$ ({\tiny{$\square$}} and {\color{red} $\circ$}, respectively) for electrical field (a) $E \perp c$ and (b) $E \parallel c$.}

\caption{(Color online) $C_p/T^3$ vs. $T$ representation of our heat capacity data. Natural $\alpha$-HgS sample: {\tiny{$\square$}}; artificial vapor phase grown sample: {\color{red} $\circ$}. The solid line represents our theoretical data (NOSO, for details see text). Earlier heat capacity data by Khattak \textit{et al.}\cite{Khattak1981} are represented by: {\color{blue}$\diamond$}.}

49 — 1007.2787

\caption{\small \capitem{a}{}Typical MEA apparatus. A tissue sample was mounted in an inverted microscope, with images projected onto it via a small video monitor at the camera port (not visible). \capcolor{Clockwise from left,} 1: suction; 2: tissue hold-down ring; 3: perfusion inflow, with temperature control; 4: preamplifier; 5: location of the multi-electrode array. \capitem{b}{} Example of a single-spike event. Each subpanel shows the time course of electrical potential ($\uVunit$) on a particular electrode in the $5\times6$ array. The electrodes are separated by $30\,\umunit$ (similar to RGC spacing). A spike from one unit is visible in the lower right corner and an axonal spike can be seen running vertically in the second column of electrodes. Data were acquired at $ 10\,\kHzunit$. After baseline subtraction and high-pass filtering (see supplement), a spatial whitening filter was applied (see \sref{pp}). \pnlabel{f:MEA}}

\caption{\small \capitem{A}{} Example of a single-spike event. Each subpanel shows the time course of electrical potential (in $\uVunit$, \capcolor{black curves}), on a particular electrode in the $5\times6$ array. After baseline subtraction and high-pass filtering, a spatial whitening filter was applied (see Methods). \capcolor{Blue curves} show the result of our fitting algorithm, in this case a single template waveform representing an individual neural unit. \capitem{B}{} Detail of a more complex event and its fit, in which a single unit fires a burst of 9 spikes of varying amplitudes (upper left channel), while a different unit fires 5 other spikes (upper right channel). \capitem{C}{} Example of an overlap event and its fit, which now is a linear superposition of 7 templates. \capitem{D}{} Detail of (C), showing signals on four of the electrodes. This time individual fit spikes are displayed. The \capcolor{red} and \capcolor{green} traces show fit templates that, although similar, differ significantly in their overall strength, and in the relative strengths of their features. The \capcolor{olive} trace shows a fit to a low-amplitude template that was later classified as unusable, and hence was discarded, by the procedure in \sref{ecr}. \pnlabel{f:expt}}

\caption{\small After fitting spikes, only noise remains. \capitem{A} Noise covariance after spatial whitening. Subpanels: spacetime covariance $\Ncov(x,y,t; x_*,y_*,t+\Delta t)$ between the central channel and its neighbors as a function of $\Delta t$, for various fixed $t$ (colored curves). Central panel (dotted line): the function $(\NfitCorMag)\exp(-\Delta t/(\NfitCorLen))$. (The various $t$ lines and the dotted line are too similar to discriminate visually.) Horizontal axes: $\Delta t$ in \msunit; Vertical axes: $\Ncov$ in $\uVunit^2$. \capitem{B} \capcolor{Blue curve,}~Semilog plot of the one point marginal probability density function of decorrelated noise samples. \capcolor{Red curve,}~same quantity, evaluated on residuals after spikes have been removed from spike events. \capcolor{Dotted curve,}~The Gaussian chosen to represent this distribution. \capitem{C} \capcolor{Green,}~detail of the same template waveform shown in \fref{template}. \capcolor{Red,}~pointwise mean of the residuals after the fit spike is subtracted from 4906 one-spike events of this type is nearly flat. This validates our assumption that spikes vary only in overall amplitude, and that noise is independent of spiking. \capcolor{Blue,}~pointwise standard deviation of the residuals, again evidence that only noise remains after fitting and subtracting spikes. \capitem{D, top} Histogram of fit values of the scale factor $A$ for a template with peak amplitude $-168\,\uVunit$ (well above noise) obtained without a prior on $A$, superposed with a Gaussian of the same mean and variance. \capitem{D, bottom} Similar histogram for a low amplitude template. A secondary bump appears, due to noise-fits, but is well separated from the main peak; a cutoff is shown as a dashed green line. The superposed Gaussian has mean and variance computed from the part of the empirical distribution lying above the cutoff. \pnlabel{f:testnoise}}

50 — 1007.3746

\caption{Measured \Leff~values as a function of Xe nuclear recoil energy. Symbols correspond to (\textcolor[rgb]{.4784,.0627,.8941}{$\lhd$})--Manzur \emph{et al.} \cite{Manzur:2009hp}; (\textcolor{red}{$\circ$})--Aprile \emph{et al.} (2009) \cite{Aprile:2008rc}; (\textcolor[rgb]{0,.5,0}{\bf$\square$})--Chepel \emph{et al.} \cite{Chepel:06}; (\textcolor{blue}{$\bigtriangleup$})--Aprile \emph{et al.} (2005) \cite{Aprile:05}; (\textcolor[rgb]{.6,0,0}{$\lozenge$})--Akimov \emph{et al.} \cite{Akimov:02}; ($\times$)--Bernabei et al.~\cite{Bernabei:01}; (\textcolor[rgb]{1.,0,1.}{$\bigtriangledown$})--Arneodo \emph{et al.} \cite{Arneodo:2000vc}. A best fit is performed according to four different methods (see text), three using the spline knots placed at energies indicated by a black `+', and one with alternate spline knots indicated by a black `$\star$'.}

\caption{Results of the three direct \Leff~measurements that extend below 10\,keV, shown alone to highlight the disagreement at low energies.}

\caption{The reported \Leff~values from Aprile\,\cite{Aprile:2008rc} (red circles) and Manzur\,\cite{Manzur:2009hp} zero field (dark purple triangles) and 1.5\,kV/cm (light purple triangles) as a function of electronic recoil equivalent energy. Superimposed are the detection efficiencies of each study. The trigger and acquisition used by Manzur differed for single phase (SP) and dual phase (DP).}

51 — 1007.4073

\caption{ (Color online) Conductance distributions $P(\ln G)$ in ordinary insulating phase {\bf B} with RBC for $L=256$ (\full) and with PBC for $L=128$ (\dashed). For ordinary insulating phase, the distribution functions of $\ln G$ are well fitted by Gaussian functions (dotted line) for any system size. For QSH insulating phase, distribution functions $P(G)$ asymptotically approach to the delta function $\delta(G-2)$. }

\caption{ (Color online) Conductance distributions at the M-QSHI transition with $L=256$. The distributions at the different points {\bf D} (\full) and {\bf D'} (\dashed) almost coincide with each other. $10^6$ samples are calculated for each critical point. }

\caption{ (Color online) Distribution functions of the largest $P(2\tau_1)$ (\full) and second-largest $P(2\tau_3)$ (\dashed) transmission eigenvalues for M-OI transitions {\bf C} with PBC and RBC. Dotted lines are the distribution functions of the conductance $P(G)$ in Fig.~\ref{fig:PG_symplectic} for $G<2$ and for $G>2$, both of which are normalized to be $1$. The latter is shifted by $-2$ along the horizontal axis to be compared with $P(2\tau_3)$. }

\caption{ (Color online) Distribution functions of the largest $P(2\tau_1)$ (\full) and second-largest $P(2\tau_3)$ (\dashed) transmission eigenvalues for M-QSHI transition {\bf D}. Dotted lines are the distribution functions of the conductance $P(G)$ in Fig.~\ref{fig:PG_M-IwithEdge} for $G<2$ and for $G>2$, both of which are normalized to be $1$. The latter is shifted by $-2$ along the horizontal axis to be compared with $P(2\tau_3)$. }

\caption{ (Color online) Distribution functions of the point-contact conductance at the upper edge, $y\_p=1$ (\full), near the edge, $y\_p=3$ (\dashed), and at a distance on the order of the localization length, $y\_p=19$ ($\cdots$), with $(p,q) =(0.750,0.146)$, for $L=80$. $10^{5}$ samples are realized. Although the two-terminal conductance is quantized for corresponding parameters, large fluctuations appear for the point-contact conductance. }

\caption{ (Color online) Fluctuations as a function of the averaged point-contact conductance in the Chalker-Coddington model (IQH system). Plots for both edge ($\square\!:\!L\!=\!80$, $+\!:\!L\!=\!160$), $y\_p\!=\!1$, and bulk ($\circ\!:\!L\!=\!80$, $\times\!:\!L\!=\!160$), $y\_p\!=\!L/2$, agree with the semi-circular relation (\full). The result for the single channel DMPK equation (Refs. \onlinecite{Gertsenshtein} and \onlinecite{Papanicolaou}) (\dashed) agrees with the semi-circle only for small averaged conductance. }

52 — 1007.5452

\caption{The magnetoresistance as a function of applied magnetic field. Here the symbols correspond to various geometries: n = 15 ({\color{Blue}{$\blacktriangledown$}}), 14 ({\color{Red}{$\diamond$}}), 13 ({\color{Green}{$\bullet$}}), 12 ({\color{Purple}{$\circ$}}), 11 ({\color{Yellow}{$\blacktriangle$}}), 10 ({\color{Cyan}{$\vartriangle$}}) and 8 ({\color{Grey}{$\triangledown$}}). The largest value of magnetoresistance seen at a field of 5T corresponds to a geometry with n = 15, with the lowest value of magnetoresistance at 5T corresponding to a geometry with n = 8.}

\caption{The magnetoresistance as a function of filling factor ($\alpha$) at 5 various applied magnetic fields: H = 0.05T ({\color{Grey}{$\triangledown$}}), 0.1T ({\color{Purple}{$\vartriangle$}}), 0.25T ({\color{Green}{$\blacktriangledown$}}), 1T ({\color{Red}{$\blacktriangle$}}) and 5T ({\color{Blue}{$\bullet$}}) }

53 — 1008.0051

\caption{\label{figch3:enf}(Color online) A plot of the fluid energy $E$ versus the dimensionless time $t/\Teddy$ (runs {\tt NSP-256A} and {\tt NSP-256B}) for Weissenberg numbers $We = 3.5$ ({\textcolor{blue}{blue circles}}) and $We = 7.1$ (black dashed line). The corresponding plot for the pure-fluid case is also shown for comparison~(\textcolor{red}{red line}). The polymers are added to the fluid at $t=27\Teddy$. }

\caption{\label{figch3:s2r}(Color online) Log-log (base 10) plots of the second-order structure function $S_2(r)$, compensated by $(r/\mathbb{L})^{-2}$, versus $r/\mathbb{L}$, for our run {\tt NSP-256B} (\textcolor{blue}{blue square}) and for the pure-fluid case (\textcolor{red}{red circle}). The regions in which the horizonal black lines overlap with the points indicate the $r^2$ scaling ranges. }

54 — 1008.0998

\caption[\cn\s model fitting.]{\cn\s model fitting. Both \subref{fig:trending:modelfit:p} and \subref{fig:trending:modelfit:p_log} show selected \pup\s and \gen\s data from (\textcolor{cyan}{$\bigstar$}) 2007 June, (\textcolor{green}{$\bigcirc$}) 2007 May, (\textcolor{grey}{$\square$}) 2005 June and (\textcolor{magenta}{$\lozenge$}) 2005 April. Models shown are (---) the HV 5-7 model, (\textcolor{blue}{-- --}) MJUO1: a modified HV model and (\textcolor{red}{-- $\cdot$ --}) MJUO3: a modified HV model that incorporates two additional layers.}

\caption[\cn\s turbulence models for MJUO.]{\cn\s turbulence models for MJUO. (---) HV 5-7 model; (\textcolor{blue}{-- --}) MJUO1: Modified HV model; (\textcolor{red}{-- $\cdot$ --}) MJUO3: Modified HV model incorporating two additional layers. MJUO3 is the recommended model.}

\caption[Wind speed analysis for observations made on 2007 May 3.]{Wind speed analysis for observations made on 2007 May 3. Models shown are (\textcolor{green}{- -}) the standard Bufton model, (---) a modified Bufton model with $H_T = 11$ km, and (\textcolor{red}{- -}) a modified Bufton model with $H_T = 11$ km and $\zeta = 20\degree$. Elevation for MJUO is 1024 m above sea level.}

\caption[Fit of \vw\s models to measured profiles.]{Fit of \vw\s models to measured profiles. Models shown are (\textcolor{magenta}{---}) MJUO1V, (\textcolor{blue}{-- --}) MJUO2V, ($\cdots$) MJUO3V and (\textcolor{red}{-- $\cdot$ --}) MJUO4V.}

55 — 1008.2140

\caption{(Left) Our target intensity profile showing the measure region (ring enclosed by \dotted) and signal region (ring enclosed by \full). (Right) The computed output of the MRAF algorithm, showing an intensity which differs from the target by $0.6\%$ in the measure region. This has been achieved by allowing unwanted light to be present outside the signal region.}

\caption{a) Initial output of MRAF algorithm. The bright cross at the bottom-right of the image is zeroth order of diffraction by the SLM. b) Light intensity after seven iterations of the feedback loop. c) The final light intensity within the ring portion of the trap (\full) is significantly smoother and more accurate than the initial output (\chain). The desired intensity distribution is shown (\dashed) for reference.}

56 — 1008.2535

\caption{(color online) Probability density function (PDF) of flagellar beat frequencies, $f$, measured from unicellular alga (\emph{C. reinhardtii}) swimming in quasi-2D liquid films ($\bar{f}=53 \pm 5$ Hz). Inset: velocity oscillations of a single swimming cell measured at 500 fps (red \textcolor{red}{\CIRCLE}), where the maximum velocity is four times the mean value (solid blue line, 134 $\mu$m/s).}

\caption{(color online) (a) The time-averaged velocity field around a swimming \emph{C. reinhardtii} (black disc) in the lab frame, where the direction of travel is to the right toward the hyperbolic stagnation point (green {\Large \textcolor{green}{$\filleddiamond$}}). Solid (red) lines are instantaneous streamlines and velocity vectors are shown on a log scale \cite{JSGLogScaling}. (b) The fluid velocity magnitude in various directions away from the cell demonstrates the predicted $u \sim r^{-1}$ scaling for a force dipole in 2D. Local minima correspond to stagnation points encountered in some directions.}

\caption{(color online) A time sequence of the velocity field evolution throughout the beat cycle of \emph{C. reinhardtii} (period $T=18.9$ ms) along with the position of the hyperbolic stagnation point (green {\Large \textcolor{green}{$\filleddiamond$}}). Insets show cell speed and beat cycle phase (lower left, see Fig. \ref{fig:4}(b) for details), and approximate flagellar shape (lower right) measured from high-speed video. (a) Early in the power stroke, the velocity field resembles a (negative) force dipole. (b) At the peak of the power stroke, the vortices lateral to the organism strengthen and sweep toward the posterior. (c-e) The vortices then shift to the anterior as the power stroke is completed. (f) At the peak of the recovery stroke, the flow field again takes the shape of a dipole, but with opposite sign. The recovery stroke velocity field (f) is weaker than the forward stroke, but is enhanced by the log scaling used in all panels \cite{JSGLogScaling}.}

57 — 1008.2984

\caption{$P(\nu_{e}\rightarrow \nu_{\tau})$ (left) and $P(\nu_{\mu}\rightarrow \nu_{\tau})$ (right) with $L=8000$~km for $\sin^{2}\theta_{13}=0.005$ ({\color{red}{solid}}) and 0.01 ({\color{green}{dashed}}), and a normal neutrino mass hierarchy.}

\caption{The CC~$\nu_{\tau}$ cross section for an isoscalar target calculated according to different models: Paschos et al. ({\color{green}{long-dashed}}), Kretzer et al. ({\color{red}{solid}}) and Hagiwara et al. ({\color{blue}{short-dashed}}). The Nuance ({\color{violet}{dotted}}) cross section prediction for an argon target is also shown.}

58 — 1008.3358

\caption{Scatter plot comparing alignment based MI estimates: Kalign \cite{lassmann2005kaa} vs. MAVID \cite{bray2003mma}. The number of pairs here is about half of that shown in Figure~2. Points on the diagonal indicate agreement between the two estimates. These data were generated using the default scoring parameters. Therefore, the plot represents a proof of principle \blue{for using MI to evaluate alignments} rather than a definitive statement about the quality of the two alignment algorithms shown. \label{fig:kalign_PG_vs_mavid_PG} }

\caption{ Scatter plot comparing $K(T_{B|A})$ estimated using compression, to the single letter Shannon information $h_1(T_{B|A})$. The diagonal, $X=Y$, is a guide for the eye. Points falling below the diagonal indicate cases where $T_{B|A}$ is not independent and identically distributed, and some letters show strong correlations. The fact that $K(T_{B|A})$ is slightly larger than $h_1(T_{B|A})$ for low entropy translation strings corresponds to the initiation cost for lpaq1 compression, which is $\approx 30$ byte independently of the sequence. \blue{The plot shows $\approx 30,000$ pairs taken from all over the animal kingdom.} \label{fig:shannon-1}}

\caption{ Scatter plot comparing p-distances $d^{\rm (p)}_{AB}$ to \red{normalized} compression distances $d^{\rm(\red{NCD})}_{AB}$ obtained from $I_{\rm align}$. The figure is based on $\approx 10^5$ mtDNA pairs, selected according to criteria discussed in the main text. Different symbols correspond to different length differences $d = |N_A-N_B|$, where $N_A$ and $N_B$ are the original (non-aligned) sequence lengths.}

\caption{ Scatter plot comparing $d^{\rm (log-MI)}_{AB}$ \blue{(based on single letter Shannon entropies)} to $d^{\rm(log-det)}_{AB}$, for $5\times 10^4$ randomly chosen pairs of species.}

\caption{\blue{Pairwise comparisons between different distance measures for complete mammalian mtDNA. Compared are the abilities of $d^{(\rm type)}$ to correctly classify a large number of quartets. First, the topologies of the quartet trees obtained with two distances $d^{({\rm type}_1)}$ and $d^{({\rm type}_2)}$ are computed. The quartets with ``worst" disagreements are then looked up in the literature. Based on the literature consensus it is decided which of the two topologies is correct -- unless both are wrong, or no consensus can be arrived at due to non-existent or conflicting literature. The four numbers in the columns 3 to 6 are the number of cases in which (1) the distance measure $d^{({\rm type}_1)}$ predicted the correct topology, (2) $d^{({\rm type}_2)}$ predicted the correct topology, (3) none of them did, and (4) no decision is possible. }}

59 — 1008.4165

\caption{\label{fig:pcs_low} (Colour online) Total photoionisation cross section of helium (\Nst{S} ground state) from the ionisation threshold to 59\,eV (red\full: present work, $\times$: theoretical values of Venuti \etal \cite{bsp:venu96}, \opensquare: experimental values of Samson \etal \cite{sfa:sams94,sfa:sams02}). The inset shows the theoretical PCS of this work and experimental values of Samson \etal on a logarithmic scale for higher photon energies.}

\caption{\label{fig:err_own}(Colour online) Relative deviation of our results in velocity form (red \opensquare) and of the results of Venuti \etal \cite{bsp:venu96} (\opendiamond) in length (orange), velocity (black), and acceleration forms (blue) from the helium photoionisation cross section calculated in this work within the length formulation. }

\caption{\label{fig:err_low}(Colour online) Relative deviation from the helium photoionisation cross section calculated in this work (length form): theoretical data of Venuti \etal \cite{bsp:venu96} (orange \opendiamond, length form), Chang and Fang \cite{bsp:chan95} (yellow \opensquare, length form), as well as Ivanov and Kheifets (single point at 40\,eV, blue\opentriangle) \cite{sfa:ivan06}, the experimental values of Samson \etal \cite{sfa:sams94,sfa:sams02} (violet \opencircle), and the compiled data of Yan \etal \cite{sct:yan98} (red $\ast$). The shaded area illustrates the error range estimated by Samson \etal for their experiment ($\pm1\%$ until 48\,eV and$\pm 1.5\%$ starting from 49\,eV).}

\caption{\label{fig:err_high}(Colour online) Relative deviation from the helium photoionisation cross section calculated in this work: Floquet calculation of Ivanov and Kheifets \cite{sfa:ivan06} (blue \opentriangle), B-spline calculation of Decleva \etal \cite{sfa:decl94} (orange \opendiamond) experimental data measured by Samson \etal \cite{sfa:sams94} (violet \opencircle) or Bizau and Wuilleumier \cite{sfa:biza95} (green \opendiamond), and the compiled values by Yan \etal \cite{sct:yan98} (red $\ast$). The shaded area illustrates the error range estimated by Samson \etal for their experiment ($\pm2\%$). }

\caption{\label{fig:pcs_high}(Colour online) The analytic tail (see \Eref{eq:tail}, solid lines) that is an approximation to the photoionisation cross section of He at high energies is shown for different parameters $Z_i$ and $Z_f$ (as specified in the graph) and is compared to the experimental values of Samson \etal \cite{sfa:sams94} (violet \opencircle) and the calculation of Decleva \etal \cite{sfa:decl94} (orange \opendiamond). The dotted and dashed lines show the first-order corrections to the dipole approximation according to \Eref{eq:tail1} due to final states with S or D symmetry, respectively. The inset shows the high-energy part of the spectrum on a linear scale. }

\caption{\label{fig:tails} (Colour online) Analytical model tails (length form, red solid lines) for different two-electron systems are compared to literature data. a) $\text{H}^{-}$: Venuti and Decleva \cite{sfa:venu97} (blue \opencircle), b) $\He$: Samson \etal \cite{sfa:sams94} (violet \opencircle), c) $\text{Li}^{+}$: Verner \etal \cite{sfa:vern96} (blue line), and d) $\text{HeH}^{+}$ (parallel contribution): Saenz \cite{csm:saen03} (blue line). For the atoms (a to c) also {\it ab-initio} cross sections obtained within this work are shown (black solid lines), while for HeH$^+$ the alternative tail proposed in \cite{csm:saen03} is additionally given (purple solid line).}

60 — 1008.5127

\caption{\label{sig}\textbf{OH collisional trap loss measurement.} (a) Semi-logarithmic plot of OH trap decay rates with ({\color{red}{$\circ$}}) and without ($\bullet$) the colliding ND$_3$ beam. The decay rate due solely to cold OH-ND$_3$ collisions is $\gamma_{\text{coll}}=\gamma_{\text{on}}-\gamma_{\text{off}}$. (b) Plot of all experimental runs measuring total cross sections with ({\color{red}{$\square$}}) and without ($\blacksquare$) a polarizing electric field. Average cross sections are determined from the weighted mean of all points and errors for the given $E$-field condition. The cross-hatched regions represent one statistical standard error.}

61 — 1009.0100

\caption{\label{mmaxntot} Mass of the most massive star versus the number of stars in the cluster (for better visibility, a small random scatter was applied to the (discrete) masses). The data are collected from the literature, with the main sources Testi et al. ({\color{red} $\blacklozenge$}) and Weidner \& Kroupa ({\color{green} $\blacksquare$}). The solid line is the mean value of $m_\mathrm{max}$ depending on $n$. The dotted lines follow the $1/6$ and $5/6$ quantiles, and should confine $2/3$rd of the observed data. (Figure taken from \citealp{maschberger+clarke2008}). }

62 — 1009.0493

\caption{\label{fig:Nu} Modification of the viscoelastic flow Nusselt number normalized by the solvent Nusselt number $Nu_s$. Main graph: $Nu/Nu_s$ as a function of the Weissenberg number $We$ for $L=10$ ($\bigtriangledown$), $L=25$ ($\Box$), $L=50$ ($\bigtriangleup$), and $L=100$ ($\bigcirc$). Power law fits of $Nu/Nu_s$(\dashed): $L=50$, $\propto We^{-0.1}$; $L=100$, $\propto We^{-0.2}$ . Insert: $Nu/Nu_s$ as a function of the polymer length $L$ for $We=10$ ($\bigcirc$) and $We=40$ ($\bigtriangleup$). }

\caption{\label{fig:P} Profiles of rms of velocity and temperature fluctuations in the lower half of the computational domain. Newtonian, HTE and HTR flows are denoted by lines without symbols, closed and open symbols, respectively. $u_\text{rms}$: \solid, $\bigtriangleup$; $w_\text{rms}$: \dashed, $\Box$; $\theta_\text{rms}$: \dotted, $\bigcirc$. }

\caption{\label{fig:E} Vertical profiles of the elastic energy contribution to the Nusselt number as described in \eqnref{eq:Nu}. $-\langle\varepsilon^p_{bl}\rangle_z$: \solid, HTE; \dashed, HTE. $-\langle\varepsilon^p_{pl}\rangle_z$: $\bigcirc$, HTE; $\bigtriangleup$, HTR. Vertical heat flux $\langle w\theta\rangle_z$: \boldsolid, HTE; \dotted, Newtonian; \bolddashed, HTR. }

63 — 1009.0894

\caption{\gray\spectral data of the four sources and the model spectra. The observed GeV and TeV data are adapted from Abdo\etal\(2010) and Aharonian \etal\(2008), respectively. The model parameters are given in Table~1. The dashed and dotted lines represent the model spectra for the upper and lower limits of $\Rc$, respectively, for both sources A and C (see text in \S\ref{sec:dif-W28}).}

64 — 1009.1494

\caption{\label{fig:snap} (Color online) (a) A pseudocolor plot of the stream function $\psi$, at a representative time in the statistically steady state, with a representative Lagrangian particle track (blue squares) superimposed on it from our run~${\tt R2}$. The symbol ${\rm o}$ indicates the beginning of the trajectory and the $\times$ sign marks its end. For an animated version see the movie file at \ytubeurl{}\/; (b) log-log (base 10) plot of the energy spectrum for our run${\tt R4}$ (\textcolor{red}{line with dots}); the black line with a slope $-3.6$ is shown for reference.}

65 — 1009.1778

\caption{ $J^{PC}=1^{--}$ $c\bar{c}$ level spectrum observed by us in data from experiment (exp) \figline{0 1 0}, as predicted by the funnel-type $c\bar{c}$ potential model of Ref.~\cite{PRD72p054026} (HC) \figline{0 0 0}, and as predicted by pure HO confinement (HO) \figline{1 1 0}. Meson and baryon loops shift the $D$ states a few MeV down/up, whereas the $S$ states shift 100--200 MeV downwards. For completeness, we also indicate the levels of the sharp, low-lying meson-meson and baryon-baryon thresholds \dotline{0.7 0.7 0.7} of the channels $D\bar{D}$, $D\bar{D}^{\ast}$, $D_{s}\bar{D}_{s}$, $D^{\ast}\bar{D}^{\ast}$, $D_{s}\bar{D}_{s}^{\ast}$, $D_{s}^{\ast}\bar{D}_{s}^{\ast}$, and $\Lambda_{c}^{+}\Lambda_{c}^{-}$. }

66 — 1009.2284

\caption{Skymaps of the arrival directions of UHECR, represented by black circles ({\color{black}$\circ$}), with energy $E\geq5.7\times10^{19}\,{\rm eV}$ observed by PAO (upper panel) and AGASA (lower panel) in the equatorial coordinates plotted using the Hammer projection. The solid red lines are the boundaries of the sky covered by PAO and AGASA experiments. The blue crosses ({\color{blue}$\times$}) represent the locations of AGN with distance $d \leq 100\,{\rm Mpc}$ from the VCV catalog. The blue square ({\color{blue}$\blacksquare$}) shows the location of Centaurus A.}

\caption{The distributions of simulated UHECR arrival directions (2700 events, represented by red dots ({\color{red}.})) obtained from the simple AGN model with $f=1$ (upper panel) and $f=0.45$ (lower panel) in PAO case. Mapping and other symbols are same as in FIG.~\protect\ref{skymap-all}.}

67 — 1009.3078

\caption{Results on the synthetic data for \ABTCone \and\ABTCtwo, with a group of asymmetric factor $k$s. As the baseline, the results for AdaBoost are also shown in these figures. ($1$) and ($2$) demonstrate decision boundaries learned by \ABTCone \and\ABTCtwo, with $k$ is $2.0$ or $3.0$. The {\color{red}$\times$}'s and {\color{blue}$\square$}'s stand for training negatives and training positives respectively. ($3$) and ($4$) demonstrate false rates (FR), false positive rates (FPR) and false negative rates (FNR) on test set with a group of $k$s ($1.2, 1.4, 1.6, 1.8, 2.0, 2.2, 2.4, 2.6, 2.8$ or $3.0$), and the corresponding rates for AdaBoost is shown as dashed lines.}

68 — 1009.3395

\caption[]{\small The higher charmonium vector states ({\color{red}exp}) as extracted by us from data: (i) the $\psi (3D)$ \cite{PRL105p102001}, in BABAR data \cite{PRL95p142001} on $e^{+}e^{-}\to J/\psi\pi^{+}\pi^{-}$; (ii) the $\psi (5S)$ and $\psi (4D)$ \cite{ARXIV10053490}, in data obtained by the Belle Collaboration on $e^{+}e^{-}\to\Lambda_{c}^{+}\Lambda_{c}^{-}$ \cite{PRL101p172001}, $D^{0}D^{\ast -}\pi^{+}$ \cite{PRD80p091101}, and $D^{0}D^{\ast -}\pi^{+}$ \cite{PRL100p062001}, as well as in the missing signal of Ref.~\cite{PRL95p142001}, and in further BABAR data on $D^{\ast}\bar{D}^{\ast}$ \cite{PRD79p092001}; (iii) the $\psi(5S)$, $\psi(4D)$, $\psi(6S)$, and $\psi(5D)$ \cite{EPL85p61002}, in the data of Ref.~\cite{PRL101p172001}; (iv) the $\psi(3D)$, $\psi(5S)$, $\psi(4D)$, $\psi(6S)$, and $\psi(5D)$ \cite{ARXIV09044351}, in new, preliminary BABAR data \cite{ARXIV08081543} on $e^{+}e^{-}\to J/\psi\pi^{+}\pi^{-}$; (v) the $\psi(7S)$, $\psi(6D)$, and $\psi(8S)$ \cite{ARXIV10044368}, in data from BABAR on $D^{\ast}\bar{D}^{\ast}$ \cite{PRD79p092001}. We also indicate the level scheme as predicted by pure HO confinement (HO) \protect\figline{0 1 0}. Meson and baryon loops shift the $D$ states a few MeV down/up, whereas the $S$ states shift 100--200 MeV downwards. For completeness, we also indicate the levels of the sharp, low-lying meson-meson and baryon-baryon thresholds \protect\figline{0.7 0.7 0.7} of the channels $D\bar{D}$, $D\bar{D}^{\ast}$, $D_{s}\bar{D}_{s}$, $D^{\ast}\bar{D}^{\ast}$, $D_{s}\bar{D}_{s}^{\ast}$, $D_{s}^{\ast}\bar{D}_{s}^{\ast}$, and $\Lambda_{c}^{+}\Lambda_{c}^{-}$. }

69 — 1009.4108

\caption{Anisotropy of the effective g-factor at different frequencies (1~GHz 14~K \fulldiamond, 9.4 GHz 15.5~K \textcolor{red}{\fulltriangle}, 34~GHz 15.5~K \textcolor[rgb]{0,0.5,0}{\fullsquare}). $\Theta$ denotes the angle between the field $H$ and the $c$-axis. Solid lines represent fits a with a uniaxial symmetry (see equation (\ref{gunixial})). The 1~GHz data are plotted at a temperature where $g(\Theta)$ is phase-shifted by $90^\circ$ compared to the behaviour above 15~K.}

\caption{Temperature dependence of the effective ESR $g$-factors for a) $H\| c$ and b) $H\bot c$ at 1 \fulldiamond, 9.4 \textcolor{red}{\fulltriangle} and 34~GHz \textcolor[rgb]{0,0.5,0}{\fullsquare}. Dashed lines indicate an extrapolation towards the 9.4 GHz data. Data points for 34~GHz at 15.5~K $H\|c$ \textcolor[rgb]{0,0.5,0}{\opensquare} and for 1~GHz at 14~K $H\|c$ \opendiamond\were extrapolated by fitting equation (\ref{gunixial}) to the angle dependence of the $g$-factor, see \autoref{gtheta}. Inset in a) shows the isothermal magnetization as function of the applied magnetic field at 2~K for $H\bot c$ \opencircle\and$H\|c$ \textcolor{red}{\opencircle}, taken from Reference \cite{Krel08a}.}

\caption{Temperature dependence of the effective ESR $g$-factor for $H\bot c$ \textcolor{red}{\fulltriangle} and $H\|c$ \textcolor[rgb]{0,0.5,0.5}{\opentriangle} at 9.4~GHz together with calculated $g$-factors using equations (\ref{hubergs}), (\ref{hubergp}) (dashed lines) and the susceptibility data measured at 0.1~T \cite{Krel08a}.}

\caption{Temperature dependence of the ESR linewidth for a) $H\|c$ and b) $H\bot c$ at 1 \fulldiamond, 9.4 \textcolor{red}{\fulltriangle} and 34~GHz \textcolor[rgb]{0,0.5,0}{\fullsquare}.}

70 — 1010.1176

\caption{SEDs used in our model, generated from a range of templates presented in \citet{draine2007}. The 60-to-100\,\micron\colour$C$ of these SEDs ranges from $-0.5$ to 0.5, indicated by a colour gradient from red (cool, $C=-0.5$) to blue (warm, $C=0.5$). The abrupt increase in the density of SEDs which peak at $\lambda \lsim 30$\,\micron\is caused by an extrapolation from the warmest\citet{draine2007} model (with $C=0.1$) to even larger values of $C$ by adding modified blackbodies with $\beta=1.5$ and temperatures ranging from 47--98\,K. A radio component is added on, based on the FIR-radio correlations; this dominates at$\lambda \gsim 0.5$--2\,mm.}

\caption{The redshift distribution of number counts above a flux density limit. The thick solid/long-dashed/dot-dashed curves show the predictions of the best-fit $\alpha=1$/$\alpha=0$/free $\alpha$ models. The coloured bands represent the 68 and 95 per cent confidence regions. At 170\,\micron\(\textsl{ISO\/}), 250 and 500\,\micron\(BLAST) and 1.1\,mm (AzTEC), redshift histograms are shown. At 850\,\micron\(SCUBA), we show the scaled histogram of \citet{chapman2005} with Poisson error bars, as described in Section~\ref{sec:reddist}. Redshift distributions at 170 \citep{patris2003,dennefeld2005,taylor2005}, 250, 500\,\micron\\citep{chapin2011} and 1.1\,mm\citep{chapin2009b} are shown for comparison only, and are not used in the fits.}

\caption{Comparison between the colours of galaxies drawn from our best-fit evolving luminosity function (contours) with data from real surveys (symbols). The left panels are galaxies selected at 70\,micron\above 5\,mJy\citep[red symbols,][]{kartaltepe2010}. The top-left panel shows 24\,\micron\vs.\70\,\micron\and the bottom-left panel shows 160\,\micron\vs.\70\,\micron. There is an approximate flux-limit of 36\,mJy in the 160\,\micron\data, which is indicated by a green dashed line. The right panels compare our model with SCUBA 450 and 850\,\micron\follow-up of 60\,\micron\sources brighter than 5.24\,Jy\citep[blue symbols whose width and height indicate measurement errors,][]{dunne2000,dunne2001}. The top right panel shows 450\,\micron\vs.\60\,\micron, and the bottom right panel shows 850\,\micron\vs.\60\,\micron. In all panels, the contours indicate the density of sources predicted from the model, with units of $(\log \mathrm{Jy})^{-2} \, \mathrm{deg}^{-2}$. Unlike Fig.~\ref{fig:locallumfunc}, which shows that our SED library and local colour-luminosity distribution can produce the correct \emph{total numbers\/} of galaxies in several bands, this comparison validates our model for \emph{individual objects\/}.}

\caption{The density of 350\,\micron\sources brighter than 20\,mJy in the colour-redshift plane for both$\alpha=1$ (blue, solid contours) and $\alpha=0$ (red, long-dashed contours) models. The background intensity maps are linear and the contours are 1, 10, 100 and 1000\,deg$^{-2}$ per unit colour per unit redshift. The distributions marginalised over redshift and colour are shown at the right and top, respectively.}

\caption{The density of 350\,\micron\sources brighter than 20\,mJy at a particular redshift in the colour-luminosity plane for both$\alpha=1$ (blue, solid contours) and $\alpha=0$ (red, long-dashed contours) models. We show the population at two redshifts, $z=0.1$ and 1.0. The contours correspond to 0.1 and 0.5 of the maximum density. The thin short-dashed lines show the flux limits at the two redshifts. These lines are independent of luminosity below $C=-0.5$ due to our choice of SED library. The solid black line shows the local colour-luminosity relationship, $p(C|L)$ (see Section~\ref{sec:locallc}), with 1-$\sigma$ limits plotted as dotted lines. This figure demonstrates the strong bias toward cooler and lower-luminosity sources when compared to the rest-frame distribution, caused by our selection function.}

71 — 1010.1230

\caption{(color online) (a) Outline of a unit cell in~\FAP. (b) View from the \textbf{\emph{c}} axis, showing the zigzag layer structure. Only Cu and Cl atoms are shown, the organic cations have been removed for clarity. (c) View from the \textbf{\emph{a}} axis, showing the arrangement of $\rm CuCl_4^{2-}$ anions and organic cations. Different color lines stand for the different magnetic interactions. Color coding is as follows: \textcolor{magenta}{Cu}, \textcolor{green}{Cl}, \textcolor{black}{C}, \textcolor{red}{H}, \textcolor{blue}{F}, and \textcolor{cyan}{N}.}

72 — 1010.2538

\caption{The snapshot distribution of nutrient (\textperiodcentered) and tumor (\textasteriskcentered) cells calculated for $d=2$, $r_c/r_n=3$, and $r_p=0.305$. Tumor cells seem to concentrate close to empty regions -- apparently in the interior they lose the competition with normal cells.}

73 — 1010.3169

\caption{\label{fig1}(a) Optical image of a sample with four nano\-wires (NWs) (horizontal) and an enlarged SEM image of one of the NWs. The (vertical) liquid channel is the only part of the sample which is not covered with photo\-resist. (b) A schematic representation of the setup used for the measurements in liquid. There are two gates, a back-gate and a liquid-gate with applied gate voltages$V_{bg}$ and $V_{lg}$. The liquid potential is measured by a calomel reference electrode and denoted as $V_{ref}$. The NW-FETs are characterized by their small-signal conductance map $G(V_{bg},V_{ref})$, shown in (c), and by the noise obtained from the temporal dependence of the source-drain current $I_{sd}$ using fast Fourier transform. In (c) the horizontal dashed line marks the border between two regimes, the NW (upper) and contact (lower) dominated regime. Noise measurements were conducted along the three solid white lines. Short solid lines and numbers represent the slopes of the equi\-conductance lines.}

\caption{\label{fig2} The noise power spectral density of the voltage fluctuations $S_V$ obtained for a source-drain bias voltage of $V_{sd}=90$\,mV for different resistances of a NW measured (a) in air and (b) in a buffer solution (Titrisol pH 7, Merck). The dashed lines indicate a$1/f$ slope. The open symbols in (b) represent the thermal noise of the NW at $V_{sd}=0$\,V (\textcolor{blue}{$\triangle$}). The calculated thermal noise of a 6.5\,M$\Omega$ resistor at 300\,K is shown for comparison (horizontal line). Inset:$S_V$ as a function of $V_{sd}$ at $7$\,Hz, measured in air (logarithmic scale). The solid red line indicates a power law with exponent two.}

\caption{\label{fig3} $S_V$ divided by the squared source-drain voltage $V_{sd}^2$ as a function of $R$ at 10\,Hz, measured in air and in buffer solution (logarithmic scale). Dashed lines are guides to the eye. Arrows indicate the transition between different regimes.}

74 — 1010.3175

\caption{\label{fig:PD} (Color online) (a) The ($B$, $T$) phase diagram of CeCoIn$_5$ at ambient pressure constructed from our measurements of the superconducting transition ($\bullet$), the onset of FL behavior at $T_{FL}$ (\textcolor{Blue}{$\blacktriangle$}), and the change in the critical behavior at $T_{cr}$ (\textcolor{Red}{$\blacktriangledown$}). In addition, literature data are shown for the maximum of the magnetoresistance $T(\rho_{max}(B))$ ($\circ$ \cite{Paglione_Smax}) and $T_{FL}$ determined by Hall effect (\textcolor{Red}{$\square$} \cite{Singh_TFL}) and resistivity measurements (\textcolor{Orange}{$\diamond$} \cite{Ronning_Hc_QCP}). (b) Tentative ($p$, $B$, $T$) phase diagram of CeCoIn$_5$. The data from (a) have been extended by published measurements under hydrostatic \cite{Tayama_To_Bc2_p,Knebel_To_Bc2_p,Lengyel_Bc2_p,Ronning_QCP_p} and under negative, chemical pressure on CeCoIn$_{5-x}$Cd$_x$ \cite{Donath_Cd,Tokiwa_To_Cd,Nair_Cd_B}. To account for the different sample qualities and varying $T_c$ definitions the measurements have been scaled to match in the regions of overlap. $B_+$ (\textcolor{Red}{$\bullet$}), and the onset of the NFL behavior (\textcolor{BrickRed}{$\blacksquare$} \cite{Ronning_QCP_p}) are projected onto the $T=0$ plane.}

75 — 1010.3219

\caption{(color online). The structure of a multi-soliton pulse. The panels show the soliton amplitudes $A_j$, velocities $V_j$, and fractions associated with the initial condition $\psi_0(x)$ \eqnrefp{multi_soliton_solutions} in the NLSE \eqnrefp{nlse}, computed using a numerical scattering transform \cite{boffetta_osborne_jcp_1992}, as a function of spatial modulation frequency $k$. Panels (a--c) correspond to $\alpha =2$ and (d--f) to $\alpha=2.2$. Relative phases are $\Phi =0$ (\red{$+$}), $\pi/4$ (\green{$\times$}), $\pi/2$ (\blue{$\bigtriangleup$}), $3\pi/4$ (\magnta{$\square$}), $\pi$ (\cyan{$\circ$}). Soliton fraction is the ratio of the combined norm of the constituent solitons, $\sum_j 2A_j$, to the total norm $\int_{-\infty}^{\infty} \vert \psi_0(x) \vert^2 dx$ \cite{NoteAmp}. In the limit $k \rightarrow \infty$, when $\alpha = 2$, $A_j \rightarrow 1/8$ [$\sum_j 2A_j \rightarrow \int_{-\infty}^{\infty} | \psi_0(x) |^2 dx \rightarrow 1/2$], and $V_j \rightarrow \pm k/8$ \cite{kodama_hasegawa_1991, afanasjev_vysloukh_j_opt_soc_am_b_1994}.}

76 — 1010.3422

\caption{PDF of the (normalizd) components of the angular velocity, $\omega_x=$ \Cred{ $\bigcirc$}, $\omega_y=$ \Cgreen{$+$}, $\omega_z=$ \Cblue{$\square$}, the dotted curve is a Gaussian and the dashed one shows a stretched exponential with $a=0.45~(F=4)$. }

\caption{PDF of angular acceleration; it is $\vec\alpha_x=$ \Cred{ $\bigcirc$}, $\vec\alpha_y=$ \Cgreen{$+$}, $\vec\alpha_z=$ \Cblue{$\square$}, the dotted curve is a Gaussian and the dashed one shows a stretched exponential with $a=0.6~(F=7.6)$}

77 — 1010.4149

\caption{ MD simulation results and correlation for the contact angle of the LJTS fluid on a smooth surface in dependence of the temperature with reduced fluid-wall dispersive energy $\FWenergy$ values of 0.09 ($\Delta$ / ---), 0.10 (\square / -- --), 0.12 ($\bullet$ / ---) as well as 0.14 ($\nabla$ / $\cdot \cdot \cdot$). The entire range between triple point and critical temperature is shown. }

78 — 1011.0037

\caption{Electrons (left panel) and positrons (right panel) spectra for our best fit parameters, with three choices of modulation. As for the antiproton spectrum and diffuse \gray{s}, electrons and positrons spectra have not been used in the fit, therefore the lines should be interpreted as predictions from our model. We show experimental data on each quantity as well: electrons -- AMS-01~\citep{AMS2000}, ATIC-2~\citep{ATIC2008}, HESS~\citep{HESS2008,HESS2009}, positrons -- AMS-01~\citep{AMS2007}, CAPRICE~\citep{CAPRICE2000}, HEAT~\citep{HEAT2004}. The dates in the legend for the data sets give the years when the corresponding data were collected. \label{ele}}

\caption{Diffuse \gray{} spectra for $10^\circ \leq |b| \leq 20^\circ$ for our best-fit CR parameters compared with {\it Fermi}-LAT data for the same region of sky. The {\it Fermi}-LAT data along with the unidentified background and source components are taken from \citet{Abdo2009diffuse}. The red hatched area represents the systematic uncertainty in the spectrum of the diffuse emission, as given in \citet{Abdo2009diffuse}. Note that our best-fit model corresponds closely to that used in \citet{Abdo2009diffuse} to derive the unidentified background and source components. Therefore, the addition of our model to these components is valid. Note that these data have {\it not} been fitted. \label{gammas}}

79 — 1011.0147

\caption{\small $D_{\parallel}$ as a function of $\sigma_{\mathrm{rc}}^{-1}$ for $T=4.0$ (\opencircle) and $T=6.0$ (\fullcircle). The linear fit (\dashedline) is rationalized in the text.}

80 — 1011.0341

\caption{Map of the HI brightness temperature at $v_\mathrm{lsr}=15$~$\kms$ in a logarithmic grey scale. The position of the polarized filaments, as well as the position of the Upper Sco subgroup (Sco OB2\_2) today (``\textasteriskcentered'') and 5 Myr ago (``+'') are shown. The brightness temperatures of the HI shells are about 3~K stronger than the background.}

\caption{The sketch depicts the position and geometry of the HI shell and its magnetic field relative to the Local Bubble. Sco OB2\_2 is drawn at its current position. GC denotes Galactic centre.}

81 — 1011.0442

\caption{% Three-dimensional perturbation velocity, normalized by the bulk velocity, as a function of the bulk Reynolds number. % The lines connect solutions at constant streamwise wavenumber: $\alpha h=0.6$ ({\color{red}dashed}), $\alpha h=1$ ({\color{blue}chain-dotted}), $\alpha h=1.5547$ ({\color{black}solid}). The symbols ($\bullet$) are for turbulent flow\cite{uhlmann:07a}. }

82 — 1011.0737

\caption{Distribution of NVSS RMs as a function of Galactic latitude, which clearly shows the broadening of the RM distribution closer to the Galactic plane. The average RM and 1-$\sigma$ spread around the average RM are calculated for 2\degr\bins in Galactic latitude, and are shown as the red line and the blue lines on either side of the red line. Only lines of sight with$|$RM$|$ $<$ 300 rad/m$^2$ are included in this and the following figures; 3\% of the lines of sight have $|$RM$|$ $>$ 150 rad/m$^2$ and are not shown in this figure. }

\caption{Distribution of the excess \sigmaRM\as a function of Galactic latitude; circles (connected by a solid line) and squares (connected by a dashed line) indicate positive and negative Galactic latitudes respectively. The data points at Galactic latitudes of$\pm$1\degr\and$\pm$ 3\degr\are missing, their\sigmaRM\values are 117 rad/m$^2$ and 85 rad/m$^2$ resp. The plotted data points contain at least 15 RMs, and their RM distribution can be fitted by a Gaussian with a reduced $\chi^2$ $<$ 2 (filled symbols) or between 2 $<$ \chitwored\$<$ 4 (open symbols). The error bars (which are often smaller than the plot symbols) are calculated as the error in the mean RM of each latitude bin. Also shown are the uncorrected \sigmaRM\calculated from Fig.\ref{b_rm.fig} (yellow line; sampled in 2\degr\bins, and averaged over positive and negative latitudes) and the\sigmaRM\(not the excess \sigmaRM!) from Table \ref{sigmaRM.table} (red symbols). The red horizontal error bars indicate the range in Galactic latitude that is covered by these data sets. To illustrate how well the \DMinf\distribution traces the excess\sigmaRM\at$|b|$ $>$ 20\degr, I plot this distribution after scaling it by a factor of 0.37 (blue line; see Sect. \ref{Sect.: ne_model}). %The small measurement errors in RM of these data sets result in the excess \sigmaRM being very close to the measured \sigmaRM. }

\caption{Distribution of \sigmaRM\$\times$ sin$|b|$ as a function of Galactic latitude, for the data points shown in Fig. \ref{sigma_rm.fig}. The data points at $\pm$ 1\degr\and$\pm$ 3\degr\are not shown; their\sigmaRM\$\times$ sin$|b|$ are 2.0 and 4.5 resp. I fitted the model described in Sect. \ref{Sect.: ratio} to the data points at $|b|$ $>$ 20\degr\for positive and negative latitudes separately, and show the best fits (solid blue lines). To illustrate the sensitivity of the modelled\sigmaRM\to$\sigma_{\mathrm{RM,MW}}$, I also draw 2 curves on either side of the best-fitting curve to the data at negative latitudes, that only differ in $\sigma_{\mathrm{RM,MW}}$ by 1 rad/m$^2$ from the best-fitting value. (dashed blue lines) }

83 — 1011.1159

\caption{Spin interfaces for one sample of size $100 \times 100 $ with the boundary condition $(1|2)$. {\color{c1} Red} color corresponds to $S=1$; {\color{c2} Green} color corresponds to $S=2$; {\color{c3} Blue} color corresponds to $S=3$; {\color{c4} Magenta} color corresponds to $S=4$ }

84 — 1011.1741

\caption{Relative difference between free surface elevations computed according to the complete and incomplete scenarios. For more details see equation \eqref{eq:d2inf}. The horizontal axis is given in seconds, while the vertical scale is a percentage. The blue solid line (\textcolor{blue}{---}) corresponds to the measure $d_2(t)$. The black dashed line ($---$) represents $d_\infty(t)$.}

85 — 1011.2486

\caption{\label{Fig6}Dependence of $J_0$ (a), $T_\mathrm{C}^{\mathrm{MFA}}$ (b) and magnetic moments (c) on the lattice parameter in Mn$_2$TiGe. Markers in (b) are the same as in (a). Magnetic moments in (d) are $m_\mathrm{Mn}$ (\opensquare) and $m_\mathrm{Ti}$ (\opentriangle).}

86 — 1011.2524

\caption{$V_{QM}$ and the (arbitrarily normalized) zero-mode \textcolor{red}{$\psi_0$} for $k_->k_+>0$, with a regularized $\delta$.}

87 — 1011.3766

\caption{\label{fig:ehatwall}(Color online) PDF of the three direction cosines of polymer end-to-end separation vector ${\bm R}$ for polymers near the wall (left panel) and for polymers at the center of the channel (right channel). Three different values of $\Wi$ are used. Namely, $\Wi=0.1 (\circ), 1.5 (\vartriangle), 4.5 (\blacksquare)$. The data for $\Wi = 1.5$ and $4.5$ coincide on each other. The PDFs of $e_x$ and $e_y$ are respectively plotted using \textcolor{blue}{continuous line with symbols} ($P_{\rm x}$) and \textcolor{red}{dashed lines with symbols} ($P_{\rm y}$). The inset shows the PDF of $e_z$, $P_{\rm z}$. }

\caption{\label{fig:ehatstrain_wall}(Color online) PDF of $e_1$, $e_2$ and $e_3$, components of the unit vector along ${\bm R}$ along the three principal directions of strain, for polymers near the wall (left panel), and for polymers near the centerline (right panel). Three different $\Wi$ are used. Namely, $\Wi=0.1 (\circ), 1.5 (\vartriangle), 4.5 (\blacksquare)$. The data for $\Wi = 1.5$ and $4.5$ coincide on each other. The PDFs of $e_1$ and $e_2$ are respectively plotted using \textcolor{blue}{continuous line with symbols} ($P_{\rm 1}$) and \textcolor{red}{dashed line with symbols} ($P_{\rm 2}$). The inset shows the PDF of $e_3$, $P_{\rm 3}$. }

88 — 1011.3861

\caption{(color online) Lift and drag forces in granular media: (a) Schematic of the experimental setup. (b) Lift force as a function of depth for the cylinder ({\color{red} $\bullet$}), square rod ({\color{cyan} $\blacksquare$}) and half cylinder ({\color{blue} $\blacktriangle$}). Gray region indicates the depth at which forces in (c) were measured. (c) Net force on rods measured in experiment ({\color{black}$\leftarrow$}) and simulation ({\color{red}$\leftarrow$}). Forces ({\color{green}$\leftarrow$}) on the intruder surfaces were measured in simulation and are scaled by four for better visibility.}

\caption{Normal ($\sigma$), and shear ($\tau$), stress on the leading surface of the cylinder as a function of tangent angle $\alpha$ compared to the stresses on a plate with the same $\alpha$. {\color{magenta} $\blacksquare$}, {\color{red}$\blacktriangle$}, {\color{green}$\bullet$}, and {\color{blue}$\blacktriangledown$} represent $\sigma_{cylinder}$, $\sigma_{plate}$, $\tau_{cylinder}$ and $\tau_{plate}$ respectively.}

\caption{(color online) (a) The drag ($|f_x|$, blue) and lift ($f_z$, red) components of the stress on a plate as a function of $\alpha$ in granular media ({\color{blue}$\blacktriangleleft$} and {\color{red}$\blacktriangle$}) as compared to a fluid with with Re~$\ll1$ (dashed lines)~\cite{gombosi1994gaskinetic}. Dash-dot red line is granular wedge model (see {\em Force model} section).}

\caption{(color online) Flow of grains and force-balance model: (a) Flow field in the vertical ($xz$) plane for three plate (solid black line) orientations. Gray boundary indicates regions with upward flow~[18]. (b) Forces on a wedge for $\alpha<\alpha_{vp}$ [regime (VP)] and on the band for $\alpha\geq \alpha_{vp}$ [regime (P)]. (c) The weight of the upward-flow region as function of $\alpha$ calculated from simulation ({\color{red} $\blacksquare$}) and fit (black) to $W=c\sin(\alpha)$, where $c=5.7$~N. (d) The average flow velocity angle $\psi$ in the upward-flow region vs. $\alpha$ ({\color{blue}$\bullet$}). (e) Normal component of the stress on the plate $\sigma$ calculated from the model (black) and measured from simulation ({\color{red}$\blacktriangle$}). The black dashed line $\alpha=\alpha_{vp}=97^\circ$ indicates the boundary between the two regimes (VP) and (P).}

89 — 1011.3868

\caption{\label{FreeEnergyDifference} Each components of the total free energy, $\beta F_{\rm PF}$ ($\blacksquare$), $\beta \left( F_{\rm P} + F_{\rm incomp} \right)$ ($\bullet$), $\beta F_{\rm S}$ ($\blacktriangle$), and $\beta F_{\rm total}$ ($\blacklozenge$) are shown. (a) Dependence on the chain length $N$ for the case with $v=0.5$, $\chi=0.0$ and $\phi = 0.1$, and (b) the dependence on the packing ratio $v$ defined in eq.~(\ref{PackingFraction}) for the case with $\chi=0.0$, $\phi = 0.3$, and $N=100$, respectively. Note that $\beta F_{\rm int} = 0$ because $\chi = 0$ for both figures. For each component, $\beta F({\rm prolate}) - \beta F({\rm oblate})$ is shown.} \end{center} \end{figure} % Figure~\ref{FreeEnergyDifference}(a) indicates that both the conformation entropy of polymers and the bending elastic energy of the membrane tend to prefer prolate shape when the chain length is increased. On the other hand, the contribution from the translational entropy of the solvents shows a more complex behavior. In the short chain length region, this contribution once decreases and then it turns to increase when the chain length becomes longer. As a sum of these components, the total free energy difference decreases monotonically as the polymer chain length is increased, leading to the equilibrium prolate shape in the long chain region. The complex behavior of the translational entropy of the solvent is understood considering the effect of the depletion layer of the polymers near the membrane. When the polymer chain length becomes comparable to the membrane thickness (around $N \sim 3$), the width of the depletion layer is negligibly thin, and the solvent distributes almost uniformly inside the vesicle, which maximize the translational entropy of the solvent. As the chain length is increased, a clear depletion layer is formed, and the solvent distribution inside the vesicle becomes inhomogeneous, which cause a decrease of the translational entropy of the solvent molecules ({\it i.e.} an increase in the free energy). The opposite behavior of this translational entropy of the solvent in the very short chain length region ($N < 3$) is an artifact of the present phase field modeling of membrane which has a finite thickness that is the same order as the gyration radius of the polymer with $N \sim 5$. Due to the smooth density profile and the finite thickness of the membrane distribution $\varphi_M({\bf r})$, either the polymer segments or the solvent molecules must come into the membrane region to fill the vacancy. Such invading molecules are strongly repelled by the membrane and cause an increase in the free energy. As the rate of this increase is different for the prolate and the oblate shapes, it leads to the steep increase of $\beta F_{\rm S}$ for $N < 3$. To understand why the conformation entropy of polymers $\beta F_{\rm P}$ prefers prolate shape, we give a simple interpretation. Let us approximate an oblate or a prolate shape with a cylinder with diameter $x$ and height $y$. These two values are determined when the total surface area and the enclosed volume are given. These conditions lead to $x^2 y = C_1$ and $x^2 + 2 x y = C_2$, where $C_1$ and $C_2$ are constants that correspond to the total enclosed volume multiplied by $4/\pi$ and the total surface area multiplied by $2/\pi$, respectively. Solving these set of equations for given $C_1$ and $C_2$ gives 3 solutions $(x_i, y_i)$ ($i = 1, 2, 3$) where $x_3 < 0 < x_2 < x_1$. Obviously, the solution $x_3 <0$ is unphysical. The other two solutions correspond to the oblate $(x_1, y_1)$ and prolate $(x_2, y_2)$, respectively. If $x_2$ is small, we obtain up to the first order in $x_2$ that $x_2 = 2 C_1/C_2 \equiv L_{\rm pr}$ (prolate) and $y_1 = C_1/(2 C_2) \equiv L_{\rm ob}$ (oblate). Thus, the ratio between the linear dimensions of the confined region for prolate and oblate cases is $L_{\rm pr} = 4 L_{\rm ob}$. Now, we estimate the increase in the conformational free energy due to such confinement. We consider an ideal chain confined in a region of size $L_{\rm C}$. As the number of segments in a blob of size $L_{\rm C}$ is proportional to $L_{\rm C}^2$, a chain made of $N$ segments can be regarded as a linear chain of $N/L_{\rm C}^2$ blobs. Therefore, the increase in the conformational free energy per chain due to the confinement is given by % \begin{equation} \Delta F(L_{\rm C}) = k_{\rm B} T \ln 2^{N/L_{\rm C}^2} = \frac{N k_{\rm B}T}{L_{\rm C}^2} \ln 2. \end{equation} % Using the fact that the number of chains is inversely proportional to the chain length $N$ because of the constant volume fraction $\phi$ inside the vesicle, and the fact that there are two directions of the confinement in the prolate case, we can estimate the difference in the total conformational free energy ({\it i.e.} $\Delta F \times$ ({\rm number of chains})) between the prolate and the oblate cases as % \begin{equation} \frac{1}{N} \left[ 2 \Delta F(L_{\rm pr}) - \Delta F(L_{\rm ob}) \right] = - \frac{7 k_{\rm B}T \ln 2}{8 L_{\rm ob}^2} < 0. \label{Confinement} \end{equation} % Equation~(\ref{Confinement}) means that the conformational free energy prefers the prolate shape, which is consistent with the results in the long chain region in Fig.~\ref{FreeEnergyDifference}(a). In Fig.~\ref{FreeEnergyDifference}(b), we show similar data as those in Fig.~\ref{FreeEnergyDifference}(a) but for the dependence on packing ratio $v$. As $v$ becomes smaller, the polymers are more strongly confined. Above simple consideration suggests that the prolate shape will be more and more stable than the oblate one when the constraint becomes stronger. Actually, we can confirm this tendency in the behavior of the conformation entropy of polymers and total free enrgy shown in Fig.~\ref{FreeEnergyDifference}(b). % \begin{figure}[t] \begin{center} \includegraphics[width=50mm]{fig_Membrane_kai_non0.eps} \caption{\label{ProlateNonZeroChi} Distributions of the membrane and the polymers are shown for the case $N=100$, $v = 0.5$ and $\phi = 0.1$. The interaction parameter $\chi$ is (a) $0.6$ (b) $0.65$ and (c) $0.7$, respectively.} \end{center} \end{figure} Finally, we consider the case where there is a repulsive interaction between the polymer segment and the solvent molecule ($\chi > 0$). In Fig.~\ref{ProlateNonZeroChi}, we show distributions of the membrane and the polymers for the case with $N = 100$, $v = 0.5$ and $\phi = 0.1$. The interaction parameter between the polymer segment and the solvent $\chi$ is (a) $\chi = 0.6$, (b) $\chi = 0.65$ and (c) $\chi = 0.7$, respectively. Compared to the athermal case ($\chi = 0$) in Fig.~\ref{VesicleShapeChi0}, the polymers distribute inhomegeneously forming a depletion layer near the membrane. While the vesicle shows a symmetric shape for a smaller value of the $\chi$-parameter (Figs.~\ref{ProlateNonZeroChi}(a) and (b)), the membrane shape becomes asymmetric for a larger value of $\chi$ ($\chi = 0.7$ in Fig.~\ref{ProlateNonZeroChi}(c)) due to an asymmetric distribution of the polymers inside the vesicle. As a conclusion, we introduced a new field theoretic model for a vesicle that encloses polymers. With this model, we succeeded in calculating the equilibrium shape deformation of the vesicle induced by the polymers. This technique has a wide variety of extensions and applications such as the fusion and fission of the membrane by introducing the Gaussian curvature into the model. \acknowledgments The authors thank Q.Du and I.Takagi for fruitful discussions. The present study is inspired by a collaboration of one of the authors (TK) with K.N.Yaegashi, M.Imai and N.Urakami. The present study is supported by Grant-in-Aid for Scientific Research on Priority Area ``Soft Matter Physics'' from the Ministry of Education, Culture, Sports, Science, and Technology of Japan, and Global COE Program at Tohoku University. \begin{thebibliography}{99} \bibitem{Cell} \Name{B.~Alberts {\it et al.}} \Book{Molecular Biology of the Cell, 4th ed.} \Publ{Garland Science, New York} \Year{2002}. \bibitem{Banat} \Name{L.M.~Banat, R.S.~Makkar \and S.S.~Cameotra} \REVIEW{Appl. Microbiol. Biotechnol.}{53}{2000}{495}. \bibitem{DDS} \Name{T.M.~Allen, {\it et al.}} \REVIEW{Science}{303}{2004}{1814}. \bibitem{Nakaya} \Name{K.~Nakaya, M.~Imai, S.~Komura, T.~Kawakatsu \and N.~Urakami} \REVIEW{Europhys. Lett.}{71}{2005}{494}. \bibitem{MolecularModelMembrane} \Name{M.~Laradji \and P.B.S.~Kumar} \REVIEW{J. Chem. Phys.}{123}{2005}{224902}. \bibitem{Urakami} \Name{T.Kurokawa, N.Urakami, K.N.Yaegashi, M.Imai \and T.Yamamoto} \Book{private communication}. \bibitem{MeshModelMembrane} \Name{H.~Noguchi \and G.~Gompper} \REVIEW{Proc. Nat. Acad. Sci.}{102}{2005}{14159}. \bibitem{AllenTildesley} \Name{M.P.~Allen \and D.J.~Tildesley} \Book{Computer Simulation of Liquids} \Publ{Oxford University Press, Oxford} \Year{1987}. \bibitem{SCF} \Name{T.~Kawakatsu} \Book{Statistical Physics of Polymers} \Publ{Springer-Verlag, Berlin} \Year{2004}. \bibitem{MatsenSchick} \Name{M.W.~Matsen \and M.~Schick} \REVIEW{Phys. Rev. Lett.}{72}{1994}{2660}. \bibitem{Du} \Name{Q.~Du, C.C.~ Liu \and X~ Wang} \REVIEW{J. Comput. Phys.}{198}{2004}{450}. \bibitem{Campelo} \Name{F.~Campelo \and A.~Hern$\acute{\rm a}$ndez-Machado} \REVIEW{Eur. Phys. J. E}{20}{2006}{37}. \bibitem{Helfrich} \Name{W.~Helfrich} \REVIEW{Z. Naturforsch.}{28 c}{1973}{693}. \bibitem{Seifert} \Name{U.~Seifert} \REVIEW{Adv. Phys.}{46}{1997}{13}. \bibitem{Du_PrivateCommunication} \Name{The integrand of the PF free energy eq.(\ref{PF_Free_Energy}) is a square of the variation of Ginzburg-Landau (GL) free energy. As the GL free energy of an interface system is proportional to the interfacial area, its variation corresponds to the mean curvature of the interface. Thus, eq.(\ref{PF_Free_Energy}) can be identified with the Helfrich's bending energy.} \bibitem{VolumeConstraint} \Name{In the actual PF simulation, the discontinuous condition $\int_{\psi({\bf r}) > 0} d{\bf r}$ is replaced by a smoothed function $\int d{\bf r} (1/2)\left\{1 + \tanh \alpha \psi({\bf r}) \right\}$ with a sufficiently large constant $\alpha$.}

90 — 1011.4304

\caption{(Color online) The SHA predicted excitation energies for $\{n_1, n_2, n_3\}$=$\{1,0,0\}, \{0,1,0\}$ and $\{0,0,1\}$ are indicated with `+' and those for $\{2,0,0\}$ are indicated with `{\brown $\times$}'. The exact lowest four excitation energies are given by `$\diamond$'. We mark three points of interest (1) $g=0.010$, (2) $g=0.045$ and (3) $g=0.150$. Note that the SHA excitation energies corresponding to $\{2,0,0\}$ cross-over those of $\{0,0,1\}$ at $g \sim 0.05$, indicating the system transiting from one regime to another.}

91 — 1011.5052

\caption{Small part of the lightcurve in the 20--40\,keV energy band as measured with IBIS/ISGRI. Data are shown as black stars with gray bars indicating the uncertainties. Time resolution is 20\,sec. The line shows the folded pulse profile over the whole data set.}

\caption{Energy resolved pulse profiles with \xte PCA (a-c) and \xte HEXTE (d). The profiles are shown twice for clarity. In panel d) the \inte pulse profile in the 20--40\,keV energy range is shown in gray for comparison. Note that the\inte profile is shifted by hand to match the peak in the \xte profile, but that the profiles are not phase aligned.}

\caption{Parameters of the phase-resolved spectra fitted with the \texttt{cutoffpl} model and the folding energy frozen to $E_\mathrm{fold}=20.0$\,keV .\textit{a)}~Pulse profile in the 13.2-14.5\,keV energy range. The different shaded areas indicate the phase bins used to extract the spectra.\textit{b)}~Photo electric absorption column $N_\text{H}$, \textit{c)}~power law normalization, \textit{d)}~power law index $\Gamma$, \textit{e)}~black body normalization, \textit{f)}~black body temperature, \textit{g)}~iron line normalization, \textit{h)}~iron line width, and \textit{i)}~\redchi-value. }

\caption{Same as Fig.~\ref{fig:phasresparas_frozCut}, but for the \texttt{cutoffpl} model with the photon index frozen to $\Gamma = 1.15$. \textit{a)}~Pulse profile in the 13.2-14.5\,keV energy range,\textit{b)}~photo electric absorption column $N_\text{H}$, \textit{c)}~power law normalization, \textit{d)}~folding energy, \textit{e)}~black body normalization, \textit{f)}~black body temperature \textit{g)}~iron line normalization, \textit{h)}~iron line width, and \textit{i)}~\redchi-value.}

92 — 1011.5152

\caption{\label{ingaas:growthtemp} a) Typical 4.2\,K photoluminescence transitions for$\mathrm{In}_\mathrm{x}\mathrm{Ga}_\mathrm{1-x}\mathrm{As}$ nanowire samples grown at temperatures of 550-590\,$^\circ\mathrm{C}$. \fullsquare~correspond to data from this work while \fulltriangle~corresponds to data from a previous study \cite{Heiss2009} using a higher In deposition rate of 0.088\,{\AA\,s$^{-1}$}. The right axis of the graph shows the indium concentration of an idealized $\mathrm{In}_\mathrm{x}\mathrm{Ga}_\mathrm{1-x}\mathrm{As}$ material with a band gap related to the transition energy. b) Typical 4.2\,K photoluminescence spectra for$\mathrm{In}_\mathrm{x}\mathrm{Ga}_\mathrm{1-x}\mathrm{As}$ nanowire samples grown at temperatures of 550, 560 and 575\,$^\circ\mathrm{C}$ under otherwise identical conditions.}

93 — 1011.5332

\caption{Using \index{list_undefined/0}\predref{list_undefined}{0} on chat 80 wrapped into the module \const{chat}. To save space only the first of the 9 reported warnings is included.}

94 — 1011.5768

\caption{HR (left panel) and colour-distance (right panel) diagrams for 63 dwarfs ({\color{red} $\square$}), 13 subgiants ({\color{magenta} $\triangle$}), and 96 giants ({\color{blue} \Large $\circ$}) analysed in this work. Filled symbols represent stars with detected planets. The distance limit of 100~pc is also shown (dotted line).}

\caption{{\it Left and middle column panels:} [C/H] and [C/Fe] as a function of [Fe/H] for our sub-samples of dwarfs ({\color{red} $\square$}), subgiants ({\color{magenta} $\triangle$}), and giants ({\color{blue} \Large $\circ$}). Superposed symbols indicate the population classification: thick disc members ($\shortmid$), thin/thick disc (+), and one halo star ($\times$). The remaining stars are thin disc members. Planet hosts are represented by filled symbols. The Sun's position is also indicated. Each panel focus on one sub-sample, and the two others are shown in light gray. {\it Right column panels:} [C/Fe] distributions comparing stars with and without planets.}

95 — 1011.6082

\caption{(a): The Hasse diagram of the tope poset~$\mathcal{T}(\mathcal{L}(\mathcal{M}),B)$ based at the tope $B:=(-1,1,1,1,1)$, and the \textcolor{red}{subposet} $\mathfrak{V}(\boldsymbol{R})$ which is the vertex set of a symmetric cycle $\boldsymbol{R}$ in the tope graph. (b): The Hasse diagram of the tope poset $\mathcal{T}\bigl(\mathcal{L}({}_{-\{1\}}\mathcal{M}),\mathrm{T}^{(+)}\bigr)$ based at the positive tope $\mathrm{T}^{(+)}:=(1,1,1,1,1)$.}

\caption{The Hasse diagram of the tope poset~$\mathcal{T}(\mathcal{L}(\mathcal{M}),B)$ based at the tope $B:=(-1,1,1,1,1)$, and the \textcolor{red}{subposet} $\mathfrak{V}(\boldsymbol{R})$ which is the vertex set of a symmetric cycle $\boldsymbol{R}$ in the tope graph, cf.~Figure~\ref{figure:1}(a).}

\caption{The Hasse diagram of the tope poset~$\mathcal{T}(\mathcal{L}(\mathcal{M}),B)$ based at the tope $B:=(-1,1,1,1,1)$, and the \textcolor{red}{subposet} $\mathfrak{V}(\boldsymbol{R})$ which is the vertex set of a symmetric cycle $\boldsymbol{R}$ in the tope graph, cf.~Figures~\ref{figure:1}(a) and~\ref{figure:2}.}

96 — 1011.6207

\caption{Maximum relative errors in energy $\Delta E/E$ (left panel) and CPU time $T_{\text{CPU}}$ (right panel) as a function of the time step, in logarithmic scale. The integration time for this analysis has been set to $2.5\times 10^6$ time units. The comparison involved $100$ orbits whose initial conditions $\left(I_1(0),I_2(0)\right)$ have been taken around the top hyperbolic border of the $I_1=2I_2$ resonance (see Fig.~\ref{PhaseSpaceBox}). Symbols are : (red) $*$ RK4 integrator, (blue) $\bigcirc$ the 4th order Yoshida integrator and (black) $\square$ the 2nd order Yoshida integrator. Straight lines denote best linear fits (left panel).} \label{Energy} \end{figure} Moreover, the GSI correctly identifies more orbits than RK4 as time steps increase (see Fig. \ref{Perc1}). Indeed, mean MEGNO values computed by means of RK4 are wrong for regular orbits when the time step is greater than $0.1$. The percentage of well identified regular orbits even reaches zero while, at the same time, $T_{\text{4th}}$ is still beyond $50\%$. This difference is less discernible for chaotic orbits, since a small drift from the orbit and/or tangent direction does not lead to completely different behaviors. However, in the following, we will use the total percentage, $p$, that presents a summary of both results. Eventually, let us point out that the lower-order $T_{\text{2nd}}$ integrator manages to identify correctly approximately the same percentage of orbits as $T_{\text{4th}}$. \begin{figure}[t] \begin{center} \psfig{file=percentage_1_full_MEGNO_eps7d-3_tmax2.5d6.eps,width=15cm} \end{center} \caption{Percentages of correctly identified orbits with respect to time step in logarithmic scale. The comparison involve $100$ orbits whose initial conditions $\left(I_1(0),I_2(0)\right)$ have been taken around the top hyperbolic border of the $I_1=2I_2$ resonance (see Fig.~\ref{PhaseSpaceBox}). Solid lines and dashed lines represent respectively the identification of regular orbits and chaotic orbits.}

97 — 1011.6372

\caption{ The X-ray radiation efficiency $\eta$ of PWNe (solid black squares) and pulsars (solid red squares) as a function of characteristic age $\tau_c$. The data has been taken from \citet{kargaltsev08} and is based on the X-ray luminosity in the 0.5-8 keV band as measured by \chandra. The open blue squares indicate the \gray\efficiencies in the 100 MeV to 100 GeV band based on\fermi\data\citep{abdo10}. The star-like symbols indicate two recently discovered pulsars (black) and PWNe (red) which were not part of the \citet{kargaltsev08} sample: the AXP 1E1547.0- 5408 \citep[$\tau_{ch}=1447$~yr][]{vink09,camilo07}, and PSR J14003-6326 \citep[$\tau_{ch}=23.7$~kyr][]{renaud10}. For these two pulsars/PWN the reported unabsorbed fluxes were converted to a 0.5-8 keV band luminosity using the distances quoted in the papers. \label{fig:eta} }

\caption{ The period and period derivative of the pulsars in the X-ray \citep{kargaltsev08} and \gray\\citep{abdo10} samples. The symbols match those in Fig.~\ref{fig:eta}. Star-like symbols represent AXP 1E1547.0- 5408 (upper right) and PSR J14003-6326. The small dots represent pulsars drawn from the ATNF catalogue \citep{manchester05}. \label{fig:ppdot} }

\caption{ The data sample plotted as $\log L_X,\log L_\gamma$ versus the best-fit models using equation~\ref{eq:fullfit}. The symbols indicate: black squares: luminosity of young PWNe; red squares: luminosity of young pulsars; black open triangles: luminosity of old PWNe; red open triangles: luminosity of olds pulsars; blue open squares: \gray\pulsars.\label{fig:bestfit} }

98 — 1012.0164

\caption{Absolute value (a) and polar angle (b) of the order parameter $Z(t)$ as a function of time for a chain of 200 rowers for different parameter sets. The global parameters are $c=0.28$ and $\beta = \pi / 4$ and for all graphs the rowers start with the same random distribution of phases. The vertical red lines ({\color{red}\legendLn}) mark the time intervall $\Delta t = 100$ for which we calculate the time averages $\bar{A} $ and $\bar{\varPhi}$. They are shown as dots on the right-hand side of the corresponding diagrams together with errorbars indicating the respective standard deviations $\sigma_{A}$ and $ \sigma_{\varPhi}$. Golden line ({\color{g_gold}\legendLn }): $ \alpha = 0.2$ , $\epsilon = 0$, and long-range hydrodynamic interactions (l.r. HI) are used. Note for this parameter set we have added $2 \pi$ to all negative values of $\varPhi (t)$ to better represent its time dependence. Green line ({\color{g_green}\legendLn }): $ \alpha = 0.2$ and $\epsilon = - 0.3$ with l.r. HI. Magenta line ({\color{g_red}\legendLn}): $ \alpha = 0.2$ and $ \epsilon = -0.3 $ but hydrodynamic interactions are artificially restricted to nearest neighbors (n.n. HI). A metachronal wave occurs that travels in a definite direction. Blue line ({\color{g_blue}\legendLn}): same situation but $\alpha = -0.2 $. The mean phase $\bar{\varPhi}$ is reversed and the metachronal wave travels in the opposite direction. }

99 — 1012.0295

\caption{Experimental neutron reflectivities plotted as a function of the diffusion vector $q$ for: the control sample ($\star $), the specimen broken at a stress intensity factor $K_I=0.61$MPa.m$^{1/2}$ (Zone 1, {\color{Red} $\star $} ) and for the specimen broken at a stress intensity factor $K_I=0.77$MPa.m$^{1/2}$ (Zone 2, {\color{Blue} +} ). We have superimposed to the experimental curve corresponding to the control sample the theoretical Fresnel reflectivity curve (red dotted line) as well as the result of the second order Born approximation (blue dashed dotted line). The experimental results corresponding to Zone 1 and Zone 2 are fitted using Eqs.~(\ref{Rlimite}),~(\ref{r1-1}) and~(\ref{r2}). The best fits correspond to $\phi_0^I=0.348 \pm 0.003$, $\ell^I \approx 43$\AA~ and $\Lambda^I \approx 35$\AA~in Zone 1, and $\phi_0^{II}=0.567 \pm 0.003$, $\ell^{II}\approx 46$\AA~and $\Lambda^{II} \approx 23$\AA~in Zone 2. Inset: Contour lines of the fit root-mean square error in the plane ($\ell, \Lambda$) in Zone 2, showing that while the combination $\ell_{eff}=\Lambda + \ell$ is rather well pinned down by the fit, $\Lambda - \ell$ is a ``soft'' direction. The relative experimental rms error per point is $0.075$, whereas the minimum relative error achieved by the fit is $0.06$.}

100 — 1012.1252

\caption{\dt\,/\,$\Delta_{max}^*$\in\smfe:\:\textcolor[rgb]{1.00,0.00,0.00}{\Squarepipe}\\textminus\\dk\, =\, 130~K;\textcolor[rgb]{0.00,0.00,0.00}{\Circpipe}\\textminus\135~K. Solid curves correspond to MNM theory with the different$\alpha \sim 1\,/\,[\breve{g}\,N(0)]$: (1) \textminus\$\alpha$\, =\, 0.1, (2)\textminus\$\alpha$\, =\, 0.2, (3)\textminus\$\alpha$\, =\, 0.3, (4)\textminus\$\alpha$\, =\, 0.6 , (5)\textminus\$\alpha$\, =\, 1.0;$T_N\,/\,T_c$\, =\, 0.7\cite{22}.}

\caption{\dt\,/\,$\Delta_{max}^*$\in\smfe\with\dk\, = 160~K (\textcolor[rgb]{0.00,0.00,0.00}{\Circsteel}) and in YBCO film with $T_c$\, =\, 87.4~K (\textcolor[rgb]{0.00,0.00,0.50}{\Circsteel}) \cite{13} as a function of $T\,/\,T^*$\($T\,/\,T_c$\in the case of the theory). Curves 1\,$\dots$\,4\correspond to BC theory\cite{23} with different $x_0 = \mu\,/\,\Delta(0)$: 1 \textminus\$x_0$\, =\, 10.0 (BSC limit), 2\textminus\$x_0$\, =\, -2.0, 3\textminus\$x_0$\, =\, -5.0, 4\textminus\$x_0$\, =\, -10.0 (BEC limit).}

101 — 1012.2135

\caption{Regime Map: Jet dynamics in ($H/D ~ - Q^*$) space. The behavior of the jet for the three fluids is shown at four different flow rates, as the height of the fall is varied. The legend is as follows: $\blacksquare$ represents the regular coiling state (RPC), the {\color{red} $\CIRCLE$} represents the doubly periodic or quasi-periodic coiling (QPC), {\color{green} \UParrow} represents multifrequency or irregular dynamics (MFC), {\color{magenta}$\blacklozenge$} represent dripping, and {\color{blue}$\bigstar$} star represent the ``leaping shampoo" state. a) Fluid S1; for the first two flow rates, the dynamics proceeds from RPC to QPC to MFC to dripping. For the latter two flow rates, the dripping state is replaced by an onset of ``leaping shampoo" state. b) Fluid S2; the dynamics is similar to fluid S1, except the dripping state is now accessible only for the lowest flow rate. The transition boundaries between the QPC state and MFC state are shifted compared to fluid S1. c) Fluid S3; the regime map shows the same generic features as for the other two fluids. d) Schematic regime map showing different flow behaviors. The pictures in the insets show the jetting dynamics either under side-view or bottom view. }

102 — 1012.2276

\caption{Simulation of a teratoma on the embryo of \colorbox{figdr}{figure 19}. At step 4 the MOC sequence belonging to the driver cell circled in red is turned into the MOC sequence of the zygote: as a consequence the development of the whole embryo starts over from the point indicated, producing a shapeless mass of cells in the neck region, composed of differentiated cells.}

103 — 1012.2590

\caption{ \pttrig dependence of the \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred for \stdassoc compared to the yield from PYTHIA scaled by 2/3. }

\caption{(a) distribution of trigger particles \zT and (b) \pthat distribution in PYTHIA at \sNNsixtytwo from PYTHIA 8.1 for \stdassoc and \stdtrig at \sNNsixtytwo and \sNNtwohundred.}

\caption{\ptassoc dependence of \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred for \stdtrig compared to the yield from PYTHIA scaled by 2/3. The inverse slope parameters from fits of an exponential to data and to PYTHIA are given in Table 1.}

\caption{\npart dependence of the \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred for \stdtrig and \stdassoc compared to the yield from PYTHIA.}

\caption{\nridge dependence on \npart for \sNNsixtytwo and\sNNtwohundred for \stdtrig and \stdassoc.}

\caption{\nridge/\njet dependence on \npart for \sNNsixtytwo and\sNNtwohundred for \stdtrig and \stdassoc.}

104 — 1012.2862

\caption{Radio and \gray\pulse profiles for the three new MSPs. The red lines correspond to the 820\,MHz discovery pulse profiles from the GBT with the 820\,MHz flux density scale on the right. The blue lines are the{\it Fermi}\pulse profiles with the photon counts and low energy cut used for the\gray\selections listed on the left. There are 2043, 2341, and 621 photons in the\gray\profiles for PSRs\mspa, \mspb, and \mspc, respectively.}

\caption{The \gray\spectra for the three MSPs. The flux points on the curve were obtained from independent fits in each energy bin, as explained in the text. The curves represent the spectrum fit with a simple exponentially cut off power-law in the full energy range of$0.1-100$\,GeV.}

105 — 1012.2907

\caption{Data from 0-12\% central \Au collisions at \sNNtwohundred from \cite{STARRidgePaper} with \trigrange{3}{4} and \assocrangevar{2} with the \ridge, the \jlc, the \as and the location of the trigger particle labeled. Colour online.}

\caption{Data from \Cu collisions at \sNNtwohundred \cite{DeSilva:2009yy} with \pT cuts gradually applied to an untriggered analysis. In this analysis both the trigger and the associated particles have the same kinematic cuts applied. Colour online.}

\caption{The \pttrig dependence of the \jly per trigger particle for \stdtrig and \stdassoc for minimum bias \dAu, 0-60\% central \Cu, and 40-80\% central \Au collisions at \sNNtwohundred and 0-60\% central \Cu and 0-80\% central \Au collisions at \sNNsixtytwo \cite{Nattrass:2008rs} with comparisons to \py version 6.4.10 \cite{Sjostrand:2006za} tune A \cite{Field:2002vt} at \sNNsixtytwo (blue) and \sNNtwohundred (red) \cite{NattrassUserMeeting}. Colour online.}

106 — 1012.3394

\caption{Polarimetric data distribution along the interface between the LC and Loop I: crosses ({\color{Red}$\times$}, colored in red at the electronic version) represent objects observed at OPD/LNA and open circles ($\circ$) represent the data acquired from \citet{heiles2000} catalogue. The thick contour indicates the position of the annular feature proposed by \citet{egger_aschenbach1995}, and the boxes roughly delimit the position of the indicated dark clouds. The data collected at OPD are mainly concentrated along the interaction ring. }

107 — 1012.3829

\caption{\label{fig:DEvVSU} (Color online) Double occupancy $D$ (upper panel) and maximum eigenvalue $\epsilon$ of $J_F(\Sigma^*)$ (lower panel) plotted against $U$ on isotherms above, close to and below the critical point (\textcolor{red}{circles}, \textcolor{OliveGreen}{triangles} and \textcolor{blue}{squares} respectively). Saddle-node bifurcation boundaries $U_{c1}$ and $U_{c2}$ occur when $\epsilon = 1$. The solid lines are composed of dense DMFT-ED solution points.}

\caption{\label{fig:CritPoint} (Color online) Landau fit (solid lines) and DMFT-ED data (markers) in the vicinity of the critical point for the shifted double occupancy, $d=D-D_c$ (upper panel), and maximum eigenvalue $\epsilon$ of $J_F(\Sigma^*)$ (markers) and the scaled and shifted second order derivative of the Landau functional $\partial_d^2 \mathcal{L}/C + 1$ (solid lines) (lower panel), plotted against, $u = U - U_c$, on isotherms with, $T= 25.0\,$meV + (0.650, 0.600, 0.575, 0.550, 0.500)$\,$meV, for \textcolor{red}{circles}, \textcolor{OliveGreen}{triangles}, \textcolor{blue}{squares}, \textcolor{magenta}{pentagons} and \textcolor{cyan}{hexagons} respectively.}

\caption{\label{fig:IPT_DEvsU} (Color online) DMFT-IPT results for the double occupancy (upper panel) and maximum eigenvalue $\epsilon$ of $J_F(\Sigma^*)$ (lower panel) plotted against $U$ on isotherms above, close to and below the critical point (\textcolor{red}{circles}, \textcolor{OliveGreen}{triangles} and \textcolor{blue}{squares} respectively).}