2009

1 — 0901.0157

\caption[Sign pattern of octahedral rotations neighboring stacking fault]{Sign pattern of octahedral rotations on opposite sides of stacking fault: octahedra in lower layer $\mathfrak{A}$ (\textcolor{red}{$\mspace{0.2mu}\Circle\mspace{0.2mu}$}) and upper layer $\mathfrak{B}$ (\textcolor{blue}{$\Box$}) and standardized displacement $\mathbf{D}$ ($\boldsymbol{\rightarrow}$). Panels~(a) and~(b) depict the two distinct choices for sign of rotation ($+$ and $-$, respectively) of $O_\mathfrak{B}$.}

\caption[Full specification of octahedral rotations neighboring stacking fault]{Full specification of octahedral rotations on opposite sides of stacking fault: octahedra in lower layer $\mathfrak{A}$ (\textcolor{red}{$\mspace{0.2mu}\Circle\mspace{0.2mu}$}) and upper layer $\mathfrak{B}$ (\textcolor{blue}{$\Box$}) and standardized displacement $\mathbf{D}$ ($\boldsymbol{\rightarrow}$), with explicit (signed) $[\pm1,\pm1,\pm1]$ rotation axes to specify full configuration. Panels~(a) and~(b) depict different configurations, with Panel~(c) showing the application of $C_{2x}$ rotation to (a) to produce a configuration equivalent to (b).}

\caption[Planar movement of oxygen ions at AO/AO stacking faults]{Planar movement of oxygen ions on opposite sides of AO/AO stacking faults: rotation of octahedra in lower layer $\mathfrak{A}$ (\textcolor{red}{$\nearrow\mspace{-8mu}\swarrow$}, outlined arrowheads on solid grid) and upper layer $\mathfrak{B}$ (\textcolor{blue}{$\nearrow\mspace{-8mu}\swarrow$}, outlined arrowheads on dashed grid) projected onto interfacial plane and resulting movement of oxygens ($\boldsymbol{\nwarrow\mspace{-8mu}\searrow}$, filled arrowheads) within such plane. Panels depict (a) high\hyp{}energy configuration with movement of neighboring oxygen atoms directly toward each other and (b) low\hyp{}energy configuration with oxygen atoms moving past each other.}

2 — 0901.0612

\caption{Simulation results of the $1200\times4000$ LDPC code using 16-bit fixed-point (\broken,\opensquare), 24-bit fixed-point (\chain,*), and floating-point (\dotted,\opencircle). The inset shows the region where the performance begins to drop in more detail.\label{fig8_2}}

3 — 0901.0738

\caption{Temperature dependence of the magnetic (101) Bragg peak as a measure of the squared magnetic order parameter. The blue symbols were measured during cooling and the red heating. The temperature ramping rate was 2.3 K/min for \fullcircle, and 4.5 K/min for \opencircle. }

4 — 0901.1913

\caption{ Light curves of various stars folded at $P_{VA1}$ and $P_\mfc2$ to demonstrate some of the ways that \fc2 can find a period other than the fundamental of the physical system. Star classification for these stars comes from the Simbad Database, operated at CDS, Strasbourg, France (%\url {http://simbad.u-strasbg.fr}) {\bf a)} {\it HIP~109303 == AR~Lac} is an RS~CVn eclipsing binary with $P_{VA1} = 1.98318 d$ which \fc2 best-fits at half the period. Due to the narrow minima, the $H=3$ Fourier expansion at the fundamental, \{$1f$, $2f$, $3f$\}, provides a worse fit than the fit at half the period \{$2f$, $4f$, $6f$\}. However, the different depths of the two minima are clearly distinguishable, showing unambiguously that $P_{VA1}$ is the true orbital period. {\bf b)} {\it HIP~10701 == AD~Ari} has $P_{VA1} = 0.269862 d$. However, \fc2 finds a period of twice that, which appears to give a better fit to the data. This is a $\delta$~Sct type variable, which are characterized by multi-modal oscillations, so power at multiple frequencies is to be expected. {\bf c)}{\it HIP~59678 == DL~Cru} is an $\alpha$ Cyg variable. These stars have complex oscillations, so different cycles at the primary fundamental may have different apparent amplitudes. The \fc2 algorithm to $H=3$, for this data set, found a lowest \chisquared at triple the fundamental period. This splits the cycles during the observation into three sets, each of which can have a quasi-independent average amplitude adjusted by the Fourier components, allowing a slight additional \chisquared reduction through noise-fitting. {\bf d)}{\it HIP~115647 == DP~Gru} is an Algol-type eclipsing variable with narrow minima. A fit to the best $H=3$ Fourier expansion is not a good model for the shape of this lightcurve, and so the \fc2 would not be very effective in finding this period unless $H$ is increased. However, the HEP data samples only 3 minima, spread over 145 orbits, and so may not be sufficiently constraining to uniquely determine a period, regardless of the algorithm used. {\bf e)}{\it HIP~98546 == V1711~Sgr} is a W~Vir type variable. $P_{VA1} = 15.052 d$, but \fc2 gives $10.566 d$. The \fc2 fit is better in the formal sense of having a lower \chisquared with the line passing closer to the data points, but it does not look like a W Vir light curve. To add to the confusion, the General Catalog Of Variable Stars (%\url {http://www.sai.msu.su/gcvs}) gives $P = 28.556 d$, but this period is not apparent in the Hipparcos data. {\bf f)}{\it HIP~12557 == W~Tri}, a semi-regular pulsating star, has $P_{VA1} = 108 d$, which is the third-best value found by \fc2 (after $17.99 d$ and $22.68 d$). \label{figlightcurves}}

5 — 0901.2956

\caption{(Color online) Proposed dynamical atom-cavity QM. The cavity couples to only one ingoing and outgoing mode, $u_{0}^{in}$ and $u_{0}^{out}$, and it is the quantum state of this mode that is stored. The pulse shape is optimized for efficient writing and reading of the state onto and from the oscillator medium inside the cavity. A symmetric pulse shape is used, so the time-reversed output is identical to the input.\textcolor{blue}{\label{fig:model}}}

\caption{(Color online) Case 1: Cavity input (dashed) and output amplitudes (solid). The dotted line gives the oscillator amplitude. The dashed-dotted cyan line represents the coupling shape in time $g(t)$. Here $t_{0}=-5$, $T=5$, $a_{0}=1$. \textcolor{blue}{\label{fig:symmetric coupling shape}}}

\caption{(Color online) Case 2: Cavity input (dashed) and output amplitudes (solid). Other lines and parameters as in Fig (2). The inset gives the detuning shapes in time: $\Delta(t)$ and $\delta(t)$.\textcolor{blue}{\label{fig:opposite detuning shape}}}

6 — 0901.3050

\caption{Variations in the FT magnitude when varying the edge energy $E_0$ between 22.70 (\green ~line) and 22.80~\AA (\red ~line). The 2.40~\AA~ peak remains on its position, while the other peaks change somewhat. }

7 — 0901.3666

\caption{An example of a time step of the stochastic model defined in section 3 for $s=3$, ${\color{blue} \filleddiamond}=I^{(1)},$ {\color{red} \textbullet}$=I^{(2)}$, ${\color{green} \filledsquare}=I^{(3)}$, $h A_1=3, \ h A_2=1, \ h A_3=1$. Each individual $I^{(i)}$ in the parent generation give raise to $h A_i$ new individuals, each of them having a probability $Q_{ji}$ of mutating into an $I^{(j)}$ one. The parent individual has a probability $h D_i$ of degradating before the selection phase. In the selection phase $N$ individuals are extracted at random (without replacement) from the pool of individuals obtained after the reproduction phase. The event probability is associated to each arrow. Notice that the probability of any individual of surviving the selection phase is independent on the specie.}

8 — 0901.3674

\caption{\label{gaps} The minimal fermion gap $\Delta$ (\full) and the maximum energy splitting between the ground states $\delta$,~\rf{gapdeg} (\dashed) as functions of $K$ for parametrizations (i) (\opencircle), (ii) (\opensquare) and (iii) (\opendiamond) given in Table~\ref{parconfs}. The fermion gap grows linearly and the degeneracy improves with increasing $K$ for all parametrizations. The fermion gap is relatively insensitive to the vortex separation, whereas the degeneracy improves when the vortices are further apart.}

\caption{\label{holos} (a) The unitarity measure, $s(\1,\Gamma_{C_l}\Gamma_{C_l}^\dagger)$, as a function of $K$ for the three configurations given in Table \ref{parconfs}. The measure of off-diagonality, $s(|R^2|,|\Gamma_{C_l}|)$ (\full), and the total fidelity, $\bar{s}(R^2,\Gamma_{C_l})$ (\dashed), as a function of $K$ for the parametrizations (b) (i) (\opencircle), (c) (ii) (\opensquare) and (d) (iii) (\opendiamond). Based on unitarity and the energy gap behavior, we expect a stable phase in the area $0.07 \lesssim K \lesssim 0.10$ bounded by the dashed vertical lines. }

9 — 0902.0355

\caption{{\red Multiplicity distributions (number of tracks in the vertex region) for events with charm candidates used in this analysis, for copper (full red) and carbon (dashed blue) targets. The corresponding averages are also shown.}}

\caption{Average $\alpha$ as a function of $x_F$ for production by baryon ($\Sigma^-$, $p$) and meson ($\pi^\pm$) beams (a) and leading and non-leading (b{\red, c}) particles. The data points are slightly offset to avoid overlapping of the error bars. Reference $\alpha$ values of $2/3$ and $1$ are shown as dotted lines. %The points outside the frame show the The points at $x_F>1$ show the average assuming that $\alpha$ does not depend on $x_F$. }

10 — 0902.0867

\caption{Acoustic phonon along $\vec{q}_{P}$ at $T=100$~K. (a) Representative IXS spectra at momenta close to $(40\overline{1})$ at $\vec{q}_{P}$ ($a^*=3.93$) and across $\vec{q}_{P}$ on a logarithmic scale with arbitrary offset. Error bars are smaller than the symbols except for the outer data points. Solid lines below that data are (shifted) resolution functions that were summed to fit the data. (b) Phonon dispersion extracted from fits to the IXS data. The dashed line is a sinusoidal model dispersion $\omega_{ni}(\vec{q})$ that serves as a guide to the eye. {\blue The inset shows a portion of reciprocal space with the scan direction through $\vec{q}_{P}$ indicated.}}

11 — 0902.1442

\caption{The quantities analyzed in the pole (or on-shell) mass approach. \label{fig1}}{\includegraphics[width=55mm]{hwrglin.eps}%% \includegraphics[width=55mm]{hrrg07.eps}}

12 — 0902.2151

\caption{ \label{fig-a-vs-B} Shifting a Feshbach resonance with laser light. {\bf a} Real part of the scattering length as a function of magnetic field. In the absence of light ($\blacktriangle$), the pole in the scattering length occurs a 1007.4 G. With 4.2 mW of light applied the Feshbach resonance is shifted to a different magnetic field. The data were recorded at a detuning of the light frequency with respect to the bound-to-bound resonance of $\Delta_L/2\pi=+107$ MHz ({\color{blue}$\scriptstyle \blacksquare$}) and $-107$ MHz ({\color{red}$\bullet$}). {\bf b} The application of the light induces two-body loss described by the rate coefficient $K_2$ for a BEC. The observed loss when changing ${\rm Re}(a)/a_{\rm bg}$ by $\pm1$ is typically one order of magnitude slower than the loss that would be incurred when using an optical Feshbach resonance. Error bars represent one statistical standard error.}

13 — 0902.2179

\caption{$z'$ filter images of the three exoplanet hosts with {\em AstraLux} at the Calar Alto observatory 2.2\,m telescope. Total integration times are 7.68\,sec for WASP-2, 15.36\,sec for TrES-2, and 11.52\,sec for TrES-4. The image scale is linear and cuts are adapted to best depicture the secondary. North is up and East is to the left.}

\caption{$i'-z'$ color vs. spectral type derived from template spectra from \citet{pic98} convolved with the sdss $i'$ and $z'$ filter curves. Overplotted are the best values for A and B component of three objects: \textcolor{green}{$\triangle$} (WASP-2), \textcolor{red}{$\diamond$} (TrES-2), and \textcolor{blue}{$\square$} (TrES-4). The bars on the three leftmost items denote the spectral uncertainty of the B component of each of the three binaries, respectively. Since we assume a SpT from the literature for the A component no error bar is given. Asterisks ($\times$\hspace{-1.42ex}$+$) denote the positions of SDSS standard main sequence stars from \citet{smi02}. These are included to verify the accuracy of the found relation as well as provide an uncertainty estimate of colors derived.}

14 — 0902.2905

\caption{Theoretical phase shift $\phiat S/N$ for the \blue\transition with zero magnetic field and a linearly polarized probe. It takes into account the three different$F' = 7/2,\ 9/2$ and $11/2$ levels of $^1\!P_1$, spanning over 60~MHz around their average frequency (center of the plot). The phase shift is represented for equally populated $m_F$ states (solid red curve) and spin-polarized atoms in $m_F = 9/2$ or $m_F = -9/2$ states (dashed blue curve). For a 90~MHz detuning, these phase shifts are comparable and amount to a few tens of mrad with typical parameters $N=10^4$~atoms and $S = 2.8\cdot 10^3 \mu$m$^2$.}

15 — 0902.3610

\caption{The one-loop h-quark mass term by the h-gluino mass ({\Large\color{blue} $\bullet$}) and chiral symmetry breaking ({\Large\color{red}{$\times$}}) insertions.}

\caption{The two-loop $(N_f-N_h)$ h-quark mass term by the h-gluino mass ({\Large\color{blue} $\bullet$}) and chiral symmetry breaking ({\Large\color{red}{$\times$}}) insertions.}

16 — 0902.4523

\caption{Universal scaling of the Rydberg fraction in the saturated excitation regime (a) The saturated Rydberg excitation is obtained with a laser coupling strength of $\Omega$=$2\pi\times$154 kHz in dense ultracold atomic clouds with densities $n= $ [\unit[3.2$\times 10^{19}$] (\textcolor{green}{$\diamond$}), \unit[6.6$\times 10^{18}$] (\textcolor{red}{$\triangledown$}), \unit[2.8$\times 10^{18}$] (\textcolor{blue}{$\square$})] {m$^{-3}$} \cite{Heidemann2007}. (b) Scanned parameter space for the individual excitation curves depicted in the $n$ - $\Omega$ plane. (c) Saturated Rydberg fraction $f_{R}$ as a function of the dimensionless parameter $\alpha = \hbar \Omega/ C_{6}n^2$ for a three dimensional configuration ($\blacksquare$) and numerical simulations (\textcolor{green}{\Large $\bullet$}). The experimental and numerical data are fitted (solid lines) to power laws of the form $f_R\sim\alpha^{1/\delta}$ from which the critical exponents $1/\delta=0.45\pm0.01$ (exp.) and $1/\delta=0.404$ (num.) are extracted. %The accuracy of this value is due to the large range of $\alpha$ by more than three orders of magnitude. }

\caption{Universal scaling behavior of the excitation rate. The rescaled excitation rate $g_{R}$ for a three dimensional density distribution is shown for experimental data ($\blacksquare$) and the corresponding numerical simulation (\textcolor{green}{\Large $\bullet$})). A linear fit to a power law $g_{R}\sim\alpha^{\gamma}$ results in a critical exponent of $\gamma=1.25\pm0.03$ (exp) and 1.15 (num).}

17 — 0903.1862

\caption[zero field characterization]{\label{mufig} \textbf{a)} Hall mobility vs.\carrier density$n_s$. \textit{Upper inset:} Sixfold valley structure of Si(111). \textit{Lower inset:} Schematic device cross-section showing two of four contacts. The (500$\mu$m)$^2$ vacuum cavity created by contact bonding preserves the H-{Si(111)} surface and serves as the gate dielectric. \textbf{b)} $\rho$ vs.\$T$ for $n_s$ ranging from 0.65 ($\blacksquare$ at top) to $7.03\times10^{11}$cm$^{-2}$ (\textcolor[rgb]{.5,0,1}{$\blacklozenge$} on bottom). The crossover from `metallic' ($d\rho/dT >0$) to `insulating' ($d\rho/dT<0$) behavior occurs near $n_{\textsc{mit}}=0.9\times10^{11}$ cm$^{-2}$(\textcolor{blue}{$\blacktriangledown$}).}

\caption[scattering times vs. temperature]{\label{tauvstfig} \textbf{Left axis:} Reduced Hall coefficient $r_H$ measured near $B=0$ (\textcolor{red}{\rule[0.5ex]{1.5em}{0.4ex}}). $r_H=0.686$ is the lower bound predicted by the simplest model. \textbf{Right axis:} Lattice scattering rate $\tau_0^{-1}$ (\textcolor{green}{$\centerdot\centerdot\centerdot$}) and valley {drag} relaxation rate $\tau_{vv}^{-1}$ (\textcolor{blue}{\rule[0.5ex]{1.5em}{0.5pt}}) {vs.} $T$, computed from $r_H$ and $\rho_{xx}$ data via Eqs.\\eqref{dragxy} and \eqref{dragxx}. The inverse quantum lifetime $\tau_q^{-1}$ (\textcolor[rgb]{.5,0,1}{$\bigstar$}) is calculated from low-$T$ SdH oscillations. The \textcolor[rgb]{1,0,1}{\rule[0.5ex]{1ex}{1pt}\rule{0.5ex}{0pt}\rule[0.5ex]{1ex}{1pt}\rule{0.5ex}{0pt}\rule[0.5ex]{1ex}{1pt}} shows the damping rate expected for a Fermi liquid (see text).}

18 — 0903.2375

\caption{{\red The LDA electronic bandstructure for 1-, 2-, 3-layers graphene and graphite. The linear bands around the K-point near the fermi level can be seen for 1-, and 3-layers, {\it i.e.} odd number of layers.}}

\caption{{\red The density of states for 1-, 2-, 3-layers graphene and graphite. Colored lines show $l$-resolved partial DOS, $s$ (dot,red), $p$ (dash,blue), and $d$ (dot-dash,green). Black full line shows Total DOS.}}

\caption{ {\blue Real (left column) and imaginary (right column) part of the dielectric function of graphite (upper panel), and of three (second upper panel), two (second lower panel), and one (lower panel) layers of graphene.} }

19 — 0903.3570

\caption{ \label{fig-shift} (Color online) Shifting the Feshbach resonance with laser light. (a) Elastic and (b) inelastic two-body scattering properties are shown as a function of magnetic field $B$. Experimental data in the presence (\color{blue}$\bullet$\color{black}) and absence ($\circ$) of the light are compared. The light power is 11.2 mW and the frequency is $\omega_L/2\pi=382,048,158$ MHz, which is 576 MHz blue detuned from the nearest bound-to-bound transition. The solid lines in (a) and (b) show fits of Eqs.\(\ref{a-FR}) and (\ref{K2-Lorentz}), respectively, to the data. The Feshbach resonance is shifted by $\sim - 0.35$ G. At $B=1006.91$ G, we measure ${\rm Re}(a)/a_{\rm bg}-1\sim 1$ and $K_2\sim 1\times 10^{-12}$ cm$^3$/s. }

\caption{ \label{fig-systematic} (Color online) Systematic study of the loss resonances. $K_2(B)$ was measured for certain values of the laser frequency at a fixed laser power of 0.47 mW. (a) The maximum $K_2^{\rm max}$ and (b) the width $W$ were determined from a fit to Eq.\(\ref{K2-Lorentz}). The experimental data for the resonances that occur at the lower ($\circ$) and higher (\color{red}$\bullet$\color{black}) value of $B$ both agree well with the predictions of the full model Eq.\(\ref{K2}) (solid lines) which is well approximated by Eqs.\(\ref{K2-approx}) and (\ref{W-approx}) (dotted lines). }

20 — 0903.5076

\caption{G vs. $V_{sg}$ as a function of the in-plane magnetic fields perpendicular to the quantum wire, $B_{\perp}$, at 100mK. (a)$V_{tp}=0V$. $B_{\perp}=$0(thick red), 2({\tiny\dotted}), 4({\tiny\dashed}), 6({\tiny\chain}) and 8T(thick black). Inset: G vs. $V_{sg}$ at $B_{\perp}=9T$(black) along with its derivative in arbitrary unit(red). (b)$V_{tp}=-1.4V$. $B_{\perp}=$0(thick red) to 9T(thick black) in steps of 1T. }

21 — 0904.0122

\caption[Beam Delivery System]{\it Beam Delivery System {\dred(BDS)} as described in the RDR. The upper part shows the region from 2200~m to 1200~m upstream of the $e^+e^-$~IP, including the polarimeter chicane at 1800~m. The lower part shows the region from 1200~m upstream to 400~m downstream of the IP, including the upstream energy spectrometer at 700~m as well as the extraction line energy spectrometer and polarimeter around 100~m downstream of the IP located at $z=0\;$m.}

\caption[Compton Cross Section]{\it{\dred(a)} Compton differential cross section versus scattered electron energy for same ({\dred black/}red curve) and opposite ({\dred grey/}green curve) helicity configuration of laser photon and beam electron. The beam energy is 250~\GeV{} and the laser photon energy is 2.3~\eV. {\dred (b)} Compton edge energy dependence on the beam energy.}

\caption[Polarimeter Chicane Parameters]{\it Magnetic chicane parameters for the BDS Compton polarimeters. {\dred The magnet labels given in parenthesis refer to Figures~\ref{fig:upst-pol}~and~\ref{fig:downst-pol}. } }

\caption[Polarimeter Detector]{\it Schematic of a single gas tube (left) and the complete hodoscope array {\dred covering the tapered exit window} (right) as foreseen for the Cherekov detectors of {\dred both} polarimeters.}

22 — 0904.0246

\caption{\sl Photograph of a Tetratex\textregistered\sample taken with a scanning electron microscope.}

\caption{\sl (left) Light yield in photoelectrons vs lifetime of the slow scintillation component for various thicknesses of TPB coatings on 3M \texttrademark\foil; (right) Conversion efficiency (expressed as absolute yield in photoelectrons for an effective slow component of\,$3.2~\mu$s) for the various thicknesses on 3M\texttrademark\foil compared to a few measurements of TTX foils.}

23 — 0904.0831

\caption{The observed color-magnitude diagram of galaxies ($M_{Rc}^{k,e} < M^*_{R_c}+2.0$) of galaxies in the cluster main-group and galaxy groups in Abell 2390. The theoretical red-sequence \citep{1997A&A...320...41K}, adjusted for calibration systematics (see \S6.3), is overlaid in each panel as the solid line, while the dashed lines define the region used for red galaxies in the calculation of \fred. Group galaxies in each panel are plotted using the symbol corresponding to the same group in Figure \ref{A2390.map}. Large symbols mark galaxies with $M_{Rc}^{k,e}<M^*_{R_c}+1.5$, which are used for the computation of \fred. \label{A2390.map.fred}}

24 — 0904.4093

\caption{Comparison between the temporal dependence of G subjected to $\Delta $T and MW protocols (see text). Sample is a In$_{2}$O$_{3-x}$ (thickness 3{4\AA }~R$_{\square}$ =28M$\Omega$ n=4.27\textperiodcentered10$^{19}$ cm$^{-3}$). (a) Illustrating the protocol for $\Delta$T=60mK chosen to match the $\Delta$G/G$\approx$30\% produced by exposing the sample to MW (power of 25dBm at f=2.416GHz) in the MW protocol. (b) The time dependence of the excess conductance of the two protocols.}

25 — 0904.4194

\caption{% \label{fig:BLS}% (color) (a) BLS spectra for 60~nm thick \CFS\film in different magnetic fields. Peaks are assigned as Damon-Eshbach (DE) or perpendicular standing spin wave (PSSW) modes. (b) field dependence of the BLS frequencies ($\blacktriangle$ Stokes lines, \textcolor{red}{$\blacktriangledown$} anti-Stokes). (c) and (d) dependence of the BLS frequencies on the angle of incidence $\varphi$ ($H$ = \unit[200]{mT} and $d$ = \unit[60]{nm}), and the film thickness $d$ ($H$ = \unit[200]{mT}), respectively. (b--d) also show the results of numerical simulations of the BLS frequencies (solid lines). See text for details. }

\caption{% \label{f:D-Heuslers} (color)(full triangles) Experimental exchange stiffness $D$ of various \CYZ\compounds. The straight dashed lines are guide for eye for L2$_1$ and B2 ordered compounds. (empty squares) Expected $D$ values for CCFA and Co$_2$MnAl when corrections for L2$_1$ order and 0\,K are taken into account (see text). Inset: Dependence of$D$ on L2$_1$ order for Co$_2$MnSi. The straight line shows a linear fit. Data determined from our previous work \protect\cite{gai08,ham09cms}. }

26 — 0905.0138

\caption{(color online) (a) The $\lambda_{ab}^{-2}(T)$ extracted from the measured susceptibility $\chi(T)$. {\Large $\mathbf{\circ}$}- without corrections, \textcolor{green}{$\mathbf{\triangle}$}-corrected for extra signal from field-penetrated part, \textcolor{blue}{$\mathbf{\lozenge}$}-additionally corrected for magnetic permeability. Solid and dashed lines show quadratic and 2-gap BCS fits, respectively(b) The $T-$dependence of the superconducting gaps, used in two-gap BCS fit of \textcolor{green}{$\mathbf{\triangle}$}. (c) Enlarged low-temperature region, showing quadratic temperature dependence of the $\lambda_{ab\:\mathrm{L}}^{-2}$}

27 — 0905.0545

\caption{The difference $\Delta_{\rm{exp-th}}$~=~$\nu_{\rm{exp}} \, - \, \nu_{\rm{th}}$ for $\nu^{\pm}_{\rm{HF}}$ between experiment and the closest theory~=~0~MHz~\cite{Korobov:01}. The second theory \cite{Kino:03APAC} is another 300~kHz less. The experimental value of $\nu^+_{\rm{HF}}$ is shown as a solid triangle ($\blacktriangle$) and $\nu^-_{\rm{HF}}$ as an empty triangle (\textcolor{red}{$\vartriangle$}). The estimated theoretical error is 1.3~MHz~\cite{Bakalov:07} and therefore too large to be shown on the scale of these graphs. (a) Pressure dependence, where the point at 250~mbar is the average of two power dependent measurements from~\cite{Pask:2008}. The points represented by the solid circle ({\LARGE $\bullet$}) and empty circle (\textcolor{red}{{\LARGE$\circ$}}) shown at $p$~=~0~mbar are $\overline{\nu}^+_{\rm{HF}}$ and $\overline{\nu}^-_{\rm{HF}}$ respectively. (b) Power dependence measured at constant pressure $p$~=~150~mbar the average of which constitutes the point at $p$~=~150~mbar in (a).}

28 — 0905.0563

\caption{{\bf The prisoner's dilemma.} Defectors save the cost $c$ of cooperation and therefore have a fitness advantage of $c$ compared to cooperators. (A), Exemplary evolutionary trajectories. A high selection strength, i.e., a high fitness difference $c=0.1$ (purple), leads to Darwinian evolution and fast extinction of cooperators, while a small one, $c=0.001$ (green), allows for dominant effects of fluctuations and maintenance of cooperation on long time-scales. We have used $N=1000$ in both cases. (B), The dependence of the corresponding mean extinction time $T$ on the system size $N$. We show data from stochastic simulations as well as analytical results (solid lines) for $T$, starting from equal abundances of both species, for different values of $c$ (see text): $c_1=0.1$ (\textcolor{violet}{$\diamondsuit$}), $c_2=0.01$ (\textcolor{blue}{\Large $\circ$}), $c_3=0.001$ (\textcolor{darkgreen}{\scriptsize $\Box$}), and $c_4=0.0001$ (\textcolor{red}{\scriptsize $\bigtriangleup$}). The transition from the neutral to the Darwinian regime occurs at population sizes $N_e^{(1)}, N_e^{(2)}, N_e^{(3)}$, and $N_e^{(4)}$. They scale as $1/c$: $N_e\approx 2.5/c$, as is confirmed by the rescaled plot where the data collapse onto the universal scaling function $G$, shown in the inset.}

\caption{{\bf Transitions and universal scaling.} We show the rescaled mean extinction time, $T/T_e$, depending on $N/N_e$, for different transitions emerging in social dilemmas (c.f.~Fig.~\ref{fig:C}). (A), Transition from the neutral regime (1), where $T\sim N$ emerges, to the Darwinian regimes (3) ($T\sim\exp N$) as well as (4) ($T\sim \ln N$). (B), From neutral dynamics in regime (2) ($T\sim\sqrt{N}$) to the Darwinian regimes (3) ($T\sim\exp N$) and (4) ($T\sim\ln N$). (C), Transition between the two neutral regimes (1) ($T\sim N$) and (2) ($T\sim \sqrt{N}$). Analytical calculations are shown as black lines, and symbols have been obtained from stochastic simulations for large (\textcolor{red}{\scriptsize $\bigtriangleup$}), medium (\textcolor{blue}{\Large $\circ$}), and small (\textcolor{orange}{$\diamondsuit$}) values of $\Ss-\Pp$ and/or $\Tt-R$. The data collapse onto universal curves reveals the accuracy of the scaling laws. In (A), we have used $\Ss-\Pp=\Tt-\Rr\in\lbrace -0.1,-0.01,-0.001,0.001,0.01,0.1 \rbrace$, while $\Ss-\Pp\in\lbrace -0.1,-0.01,-0.001,0.001,0.01,0.1 \rbrace$, $\Tt-\Rr=1$ in (B), and $\Ss-\Pp=0$, $\Tt-\Rr\in\lbrace 0.001, 0.01, 0.1\rbrace$ in (C).}

29 — 0905.1718

\caption{Fits of the mathematical model given in \eref{S}--(\ref{eqn:R}) to data on killing of NP396- and GP276-pulsed targets by effector and memory CD8$^+$T cells. The model assumes that the rate of recruitment of targets into the spleen depends on the spleen size and that killing of peptide-pulsed targets depends on the average frequency of epitope-specific CD8$^+$T cells. Panels A-B show the recruitment of unpulsed targets into the spleen, and panels C-F show the decline in the the ratio of the frequency of peptide-pulsed to unpulsed targets over time. Panels A, C, and E are for acutely infected mice, and panels B, D, F are for LCMV-immune (memory) mice. Panels C and D are for NP396-pulsed targets and panels E and F are the GP276-pulsed targets. Black dots ($\bullet$) denote measurements from individual mice, and black lines denote the log average value per time point. Red boxes (\textcolor{red}{$\mathbf \Box$}) show the number of recruited cells predicted by the model for individual mice (panels A and B) or the predicted average ratio $R$ (panels C-F). Red lines show the log average between individually predicted values. Note the different scale for killing of target cells in acutely infected (panels C and E) and memory (panels D and F) mice. Parameters providing the best fit of the model are shown in \tref{parameters}. The lack of fit test confirms good quality fits of the data (after removing two outliers, $F_{30,162}=0.79$, $p=0.77$). Because of the reduced number of parameters, these fits of the data are only moderately worse than those obtained in our previous study \cite{Ganusov.jv08}. %$F_{30,162}=1.48$, $p=0.07$ }

30 — 0905.3100

\caption{\label{susc}(a) $M/H(T)$ of NpPd$_3$ at $H=1.1$ T ($\blacksquare$), 4 T ({\color{red} $\circ$}) and 7 T ({\color{blue} $\blacktriangle$}), showing transitions at $T = 10$ K and $30$ K. (b) Curie-Weiss fit to $H/M(T)$ for $H=1.1$ T.}

\caption{(a) $M/H(T)$ of (U$_{0.5}$Np$_{0.5}$)Pd$_3$ at $H=1.1$ T ({\color{blue} $\square$}) and 4 T ({\color{red} $\bullet$}), showing a transition at $T \sim 12$ K. (b) Curie-Weiss fit to $H/M(T)$ for $H=1.1$ T giving an effective moment of $2.956\pm0.003$ $\mu_B$/An.}

\caption{\label{rhofield} $\rho(T)$ for NpPd$_3$ in $H=0$ T ($\blacksquare$) and 9 T ({\color{red}$\circ$}). In zero field the two transitions at 10 and 30 K can be seen clearly. In 9 T the upper transition is smoothed away, while the lower transition is shifted down in temperature.}

\caption{\label{res50} (a) $\rho(T)$ of (U$_{0.5}$Np$_{0.5}$)Pd$_3$ at $H=0$ T ({\color{red}$\blacksquare$}) for $T=2-300$ K, showing a smooth change from a positive to a negative gradient at $T\sim50$ K. The transition observed in the magnetisation measurements is seen more clearly in the temperature derivative (black line). (b) Below 5 K the resistivity varies as $T^2$, in very good agreement with Fermi-Liquid Theory.}

\caption{\label{res50field}$\rho(T)$ for (U$_{0.5}$Np$_{0.5}$)Pd$_3$ in $H=0$ ({\color{red}$\blacksquare$}) and 9 ({\color{blue}$\circ$}) T. In zero field a transition can be seen at 12 K. In 9 T the transition is smoothed away.}

\caption{\label{np50cp}$C_P/T$ for (U$_{0.5}$Np$_{0.5}$)Pd$_3$ ({\color{blue}$\bullet$}) and ThPd$_3$ ($\square$) at $H=0$ T for $T=2-100$ K.}

\caption{\label{HC5} $C_P/T$ in (U$_{0.95}$Np$_{0.05}$)Pd$_3$ for $H=0$ T ({\color[rgb]{0.00,0.40,0.29}$\vartriangle$}), $H=5$ T ({\color{red}$\blacksquare$}) and $H=9$ T, ({\color{blue}$\circ$}) and in ThPd$_3$ ({\color[rgb]{0.5,0.00,0.5}{$\blacktriangledown$}}).}

\caption{\label{HC1} $C_P/T$ of (U$_{0.99}$Np$_{0.01}$)Pd$_3$ for $H=0$ T ({\color{blue}$\square$}), $H=5$ T ({\color{red}$\bullet$}) and $H=9$ T ({\color[rgb]{0.00,0.40,0.29}$\vartriangle$}), and in ThPd$_3$ ({\color[rgb]{0.5,0.00,0.5}{$\blacktriangledown$}}). The general shape is very reminiscent of that for pure UPd$_3$. The arrows show how the features associated with transitions evolve as a function of the applied field. The feature labeled with a question marked arrow may not be due to an intrinsic property of the system.}

31 — 0905.3353

\caption{By fitting the artificial data described in the text, we estimated the average turnover rate using three models: the Asymptote model (\eref{L-deboer}), the Exponential model (in which a fraction $\alpha$ of cells have exponentially distributed turnover rates, \eref{L-expa}), and the Gamma model (with gamma-distributed turnover rates, \eref{L-gamma}). Estimated mean values and 95\% confidence intervals obtained by bootstrapping the residuals with 1000 simulations are shown. Data were generated using the Gamma model ($\Box$), the Exponential model (\textcolor{light-gray}{$\blacksquare$}) and the Two-populations model (\textcolor{dark-gray}{$\blacksquare$}). Labeling periods were 7 (panel A) and 15 (panel B) days. Horizontal dashed lines denote the actual average rate of lymphocyte turnover in all data, $\pp=0.1$/day. Note that in this example, the Asymptote model always underestimated the average rate of cell turnover, and that there is a systematic 2-fold underestimation of the average turnover by all models when the data from the Two-populations model were fitted. This is because all three models fail to describe the relatively rapid accumulation of the label at early time points (see \fref{simdata}E--F). }

\caption{Estimates of the average turnover rates of CD4$^+$ (panel A) and CD8$^+$ (panel B) T cells in four healthy humans obtained by fitting with three different models: the Asymptote model ($\Box$, the Exponential model (\textcolor{light-gray}{$\blacksquare$}), in which a fraction $\al$ of the cells have exponentially-distributed turnover rates, and the Gamma model (\textcolor{dark-gray}{$\blacksquare$}) with gamma-distributed turnover rates. Best fits of the data are shown in \fref{mohri.controls}, and estimates of all parameters of the models are shown in Appendix (\tref{mohri.controls.cd4} and \ref{tab:mohri.controls.cd8}). Confidence intervals were obtained by bootstrapping the residuals with 1000 simulations. Horizontal dashed lines denote the mean of the estimated turnover rates in all patients and all models for CD4$^+$ ($\pp = 0.46\%$ per day) and CD8$^+$ ($\pp = 0.29\%$ per day) T cells. Note that all models deliver very similar estimates for the average turnover rate $\pp$, with the exception of the estimated CD8$^+$ T-cell turnover rates in individual c1 which are highly model-dependent.}

32 — 0906.2596

\caption{Vacancy levels $E_{\Gamma_8}$ (\opendiamond) and $E_{\Gamma_7}$ (\opensquare) for $\lambda_{\rm Si}=43$ meV together with $E_{T_2}$ (\opencircle) for $\lambda_{\rm Si}=0$ as functions of the inverse of total number of $\rm{\bf k}$-points $1/N_{k}$.}

33 — 0906.4276

\caption{Dependence of the intrinsic energy change $\Delta\expect{\HO_{\intr}}_{\beta}$ (a) and of the expectation value $\expect{\HO_{\cm}}_{\beta}$ (b) on $\beta$ in IT-NCSM(seq) (\symbolcircle[FGBlue]) and IT-CI($4p4h$) (\symboldiamond[FGRed]) calculations for the ground state of \elem{O}{16}. In IT-CI($4p4h$), $e_{\max}=5$ and $\hbar\Omega=30$ MeV. In IT-NCSM(seq), $N_{\max}=8$ and $\hbar\Omega=30$ MeV.}

34 — 0906.4616

\caption{(a, {\color{blue} Media 1}) Experimentally recorded images of the light scattered from a single nematicon at different positions $\Delta y$ in the cell relative to the upper plate; the input beam power is 4.7~mW. (b) Averaged distance (solid line) at which the intensity of the scattered radiation is at its half-maximum; data points are shown by colored dots in black for a single nematicon, in red for two in-phase and in blue for two out-of-phase nematicons.}

\caption{Typical intensity distributions for the interacting in-phase (a, b) and out-of-phase (c, d) nematicons. The input power of each beam is 2.35~mW. The profiles (a-d) recorded for the beams in the middle of the cell, $\Delta y=32.5\,\mu$m; (b) at $z=0.85$~mm and (d) at $z=1.2$~mm. Similar to Fig.~\ref{fig2} we also compare the profiles for both the in-phase (e, {\color{blue} Media 2}) and out-of-phase (f, {\color{blue} Media 3}) nematicons versus their position in the cell $\Delta y$.}

\caption{Experimental results on the interaction between two nematicons. (a, {\color{blue} Media 4}) Effects due to varying the $z-$position of the focal plane of the input objective lens (5.5~mm) with respect to the input facet; here $\Delta z$ is the distance between the objective and the NLC cell. (b, {\color{blue} Media 5}) Effects due to variations in the launch angle $\alpha$ between two in-phase nematicons.}

35 — 0907.0293

\caption{Preliminary adaptively smoothed (s.n.r. = 10) LAT counts map of a $10^\circ \times 10^\circ$ region centred on the LMC for the energy range 200 MeV -- 100 GeV. Contours show the extinction map from \cite{Schlegel1998} as an approximate tracer of the total gas column density in the LMC. Ten linearly spaced contour levels are plotted. The diamond in the north-east of the image designates the location of the blazar CRATES J060106-703606 \cite{Healey2007} that contributes at a low level to the \gray{} emission in this area. }

36 — 0907.0435

\caption{{\it Top:} Optical spectrum of J0303-0019 (black histogram), the same spectrum smoothed over 20\,\AA\(thick grey curve) and the noise per pixel (blue line). A simultaneous fit (thick green curve) of the \nv\line and the broad and narrow components of the\lya\line is over plotted on top of a power-law with slope of$\alpha_\lambda = -1.5$, normalized to a line free region of the spectrum (short-dashed line). In addition, the wavelength of the onset of the GP trough and of \lya\at the systemic redshift are indicated by vertical dashed lines. The normalized SDSS QSO composite spectrum by\citet{vdb01} is over plotted in red for a visual comparison of the line widths. {\it Middle and bottom:} NIR $Y$ and $K$ band spectra of J0303-0019, including the C\,IV and Mg\,II lines, respectively, smoothed over 5 pixels. The grey regions are seriously affected by sky lines and omitted from the fit. Lines and colors as in top panel. The slope of the power-law continuum here is$\alpha_\lambda = -1.0$. \label{fig:spectra}}

\caption{The accretion luminosity of J0303-0019 and J0005-0006: the figure shows the M$_{\rm BH}$ -- L$_{\rm bol}$ diagram reproduced from S08, including $\sim$60,000 SDSS QSOs at $0.1 \le z \le 4.5$. Colors show M$_{\rm BH}$ estimates using the following virial estimators: red for H$\beta$, green for Mg\,II, blue for C\,IV. Three diagonal lines show 0.01, 0.1, and 1$L_{\rm Edd}$, as indicated. Superposed are the data points for J0303-0019 (this work) and for J0005-0006 (K07). Here, we have applied the cosmology specified in S08, resulting in M$_{\rm BH}$ (L$_{\rm bol}$) larger by 5\% (7\%), but \redd\similar to those in our cosmology.\label{fig:mass_lum}}

37 — 0907.1012

\caption{The Performance Comparison of The Five Algorithms of ACO, ACO-K-Means, ACO-SLC ACO-SLC-LWCR and ACO-SLC-Mixture: The figure shows that ACO-SLC algorithm, ACO-SLC, ACO-SLC-LWCR and ACO-SLC-Mixture are faster than ACO by 415\symbol{126}10736, 390\symbol{126}10192 and 257-9419 of factors respectively! However, some solutions of ACO-K-Means and ACO-SLC have low quality. ACO-SLC-Mixture can process mixture distribution and its inaccuracy ratio is less than ACO in most cases, and is bigger than ACO by 2\% at most. }

38 — 0907.1160

\caption[]{Schematic figure of the Landau levels as a function of magnetic field. The energy scale of the onset temperature of the first increase of resistivity, $T_0$, is located between the Zeeman gap $2E_Z$ and the $E_1 -E_0$ gap, where the Boltzmann factor $k_{\rm B}$ is taken as unit. {\blue The energy scale of the second increase of resistivity at low temperatures, $T_l$, is smaller than the Zeeman gap.} }

\caption[]{ The $n$-dependences of $I_{0,j}/I_{0,0}$ with ${\bf R}_j =(na,0)$ (the closed circles) and ${\bf R}_i =(0,nb)$ (the open circles), where we take $b=\sqrt{2}a$. The dashed and dotted lines are {\red a guide to the eye}. }

39 — 0907.1184

\caption{The optimum signal-to-noise ratio for each time delay $T$ at $p = 250$~mbar. The circles ({\LARGE $\bullet$}) represent the experimental results from Pask \emph{et.~al.}~\cite{Pask:2008} while the triangles (\textcolor{red}{$\blacktriangle$}) represent the simulated data. The larger signal results from $\Gamma_{i} = 2.3 \times 10^5$~s$^{-1}$ and the smaller from $\Gamma_{i} = 4.5 \times 10^5$~s$^{-1}$}

40 — 0907.1461

\caption{(color online) (a) DMRG results for the magnetization of a $S=1$ ladder with $N=39$ rungs as a function of $J_{\parallel} + J_{\times}$ for $r=J_{\parallel}/J_{\times} = 0.9$. (b) Phase diagram as obtained from the DMRG results for this finite system. The notation $[i:j]$ refers to the value of the total spin of two consecutive rungs $i$ and $j$ as described in the text. Thick solid lines indicate the position of jumps in the magnetization curves. (c) Value of $\langle S_z \rangle$ for the sites of two consecutive rungs at the center of the system as a function of $J_{\parallel} + J_{\times}$ inside the plateaux at $M = 1/2$ (\textcolor{blue}{$\bigcirc$}), $M = 1$ (\textcolor{red}{$\bigtriangleup$}) and $M = 3/2$ (\textcolor{green}{$\square$}).}

\caption{(color online) (a) DMRG results for the magnetization of a $S=3/2$ ladder with $N=39$ rungs as a function of $J_{\parallel} + J_{\times}$ for $r=J_{\parallel}/J_{\times} = 0.9$. (b) Phase diagram as obtained from the DMRG results for these finite systems. The notation $[i:j]$ denotes the value of the total spin of two consecutive rungs $i$ and $j$ as described in the text. Thick solid lines indicate the position of jumps in the magnetization curves. (c) Value of $\langle S_z \rangle$ on the sites on two consecutive rungs at the center of the system as a function of $J_{\parallel} + J_{\times}$ inside the plateaux at $M = 1/2$ (\textcolor{cyan}{$\bigcirc$}), $M = 1$ (\textcolor{red}{$\times$}), $M = 3/2$ (\textcolor{blue}{$\square$}), $M = 2$ (\textcolor{green}{$\ast$}) and $M = 5/2$ (\textcolor{magenta}{$\bigtriangleup$}).}

41 — 0907.1593

\caption{(a) Illustration of the Hubbard square\,; (b)-(d) representations of the three different types of configurations present in the ground state of the half-filled Hubbard square. The symbols stand for:$\opencircle$ empty site, $\fullcircle\!\!\!\full\!\fullcircle$ singlet bond, $\boldsymbol{\uparrow}\!\boldsymbol{\downarrow}$ doubly occupied site.}

\caption{Comparison of superconducting order parameters for the simple Hubbard model and four different approaches: full Gutzwiller projection for the $t-J$ model ($U=12\, t,\, \fullsquare$---$\fullsquare$), partial Gutzwiller projection ($U=10\, t,\, {\color{red}\fulltriangle}\!${\color{red}---}$\!{\color{red}\fulltriangle}$), our ansatz ($U=8\, t,\, {\color{green}\fullcircle}\!\!\!${\color{green}---}$\!{\color{green}\fullcircle}$), and Gaussian Monte Carlo ($U=6\, t,\, {\color{blue}\fulldiamond}$).}

\caption{Superconducting order parameter as a function of doping for the Hubbard model including next-nearest-neighbour hopping: ${\color{red}\fullcircle}\!\!\!${\color{red}---}$\!{\color{red}\fullcircle}$\; our variational data,$\opentriangle\!$---$\!\opentriangle$\;results obtained with the Quantum Cluster method. All data points were obtained using the parameter values\,$U=8\, t,\; t'=-0.3\, t$.}

42 — 0907.1907

\caption{Comparison of Gramians computed using (a) ERA, (b) balanced POD, and (c) ERA with pseudo-adjoint modes: The empirical Hankel singular values ( $\solid$) and the diagonal elements of the controllability ({\color{red}$\dashed$, $\circ$}) and observability ({\color{blue}$\chndot$, $\times$}) Gramians with different order of modes (e.g., 4, 10, 20) in output projection. }

\caption{$H_2-$norm of the error with increasing order of the reduced-order models: exact output of the output-projected system~($\dashed$); models obtained using balanced POD~({\color{black}$\dashed$,~$\square$}), ERA({\color{blue}$\solid$,~$\times$}), ERA with pseudo-adjoint modes({\color{green}$\solid$,~$\triangledown$}) , and POD({\color{red}$\solid$,$\triangle$}). A 20-mode output projection is used in ERA, balanced POD, and ERA with pseudo-adjoint modes. }

\caption{The first output, output~$a_1$, from the impulse-response simulation: results of full-simulation ({\color{black}$\solid$,~$\circ$}), compared with those of $16$-mode reduced order model by ERA({\color{blue}$\dashed$,~$\times$}), $30$-mode model by ERA with pseudo-adjoint modes({\color{green}$\chndot$ ,~$\triangledown$}) , and $30$-mode model by POD({\color{red}$\dashed$,$\triangle$}) . A $20$-mode output projection is used in ERA and ERA with pseudo-adjoint modes.}

\caption{Singular-value plots: The full system~({\color{red}$\chndot$}) and 30-mode models obtained using balanced POD~($\dotted$), ERA~($\solid$), ERA with pseudo-adjoint modes({\color{green}$\dashed$}), and POD({\color{blue}$\chndot$}), all with a 20-mode output projection. ERA and balanced POD models generate almost identical plots. }

43 — 0907.2949

\caption{Example of two situations that are \blue{indistinguishable} by the nodes.}

44 — 0907.3313

\caption{The figure shows $\bar n_{m}$ (\textcolor{blue}{$\bullet$}) and $\Gamma_m=\Gamma_m^T+\Gamma_{opt}$ ($\bullet$) versus $\bar n_p$. The solid black curve is a fit to the measured $\Gamma_m$. The solid blue curve is the expected value of $\bar n_{m}$ assuming ideal values of $\bar n_{SR}$ and $\dot n_T=3\cdot 10^4$ quanta/sec.}

45 — 0907.4999

\caption{{\bf Non-equilibrium phase diagrams delineating observable shear thickening regions in terms of the associated stress range.} Stress range as a function of applied magnetic field $B$ (a), applied electric field $E$ (b), and packing fraction $\phi$ (c). The boundaries of the shear thickening regime are set by the local minima ($\blacktriangledown$) and maxima ($\blacktriangle$) of the viscosity curves in Figs.~\ref{fig:STwfield} and \ref{fig:confinement}. ({\color{Cyan} $\bullet$}): the yield stress $\tau_y$, below which suspensions are jammed. ({\color{Red} $\bullet$}): the predicted onset of shear thickening $\tau_m$ evaluated from Eq.~\ref{eqn:tau_min} at the measured $\dot\gamma_{m}$, demonstrating that the boundary is determined by the total shear thinning stress $\tau_{HB}$, regardless of the source of the yield stress. For panels a and b the values of $\epsilon$ used is that measured for zero attractions, showing that the shear thickening stress term is independent of field strength. For panel c the value of $\epsilon = 0$ is used which is measured at the highest packing fractions where shear thickening can be observed, showing that the phase boundary is equal to the shear thinning stress $\tau_{HB}$ in the limit of $\epsilon=0$. Solid black lines: the measured stresses at the upper and lower phase boundaries in the limit of zero field and small $\phi$. These coincide with the measured phase boundaries for $B=0$ and $E=0$. Solid blue line: prediction of the ER yield stress from two-particle interaction (see Suppl. Mat.). Dotted blue lines: guides to the eye for the phase boundary between shear thinning and jammed regimes. ({\color{Red} $\circ$}): predicted $\tau_m$ in cases where no shear thickening regime was found using model predictions for $\dot\gamma_m$. In each case, these values are close to or above the upper stress boundary, showing that the reason shear thickening was not found was because $\tau_m$ exceeded the shear thickening stress range. }

\caption{Phase diagram showing different approximations of the shear thickening phase boundary. Lower ($\blacktriangledown$) and upper ($\blacktriangle$) boundaries of the shear thickening regime are as in Fig.~4\label{fig:phasediagrams}a. Purple circles ({\color{Purple} $\bullet$}): estimate in the limit of $\epsilon=0$ giving $\tau_m = \tau_{HB}(\dot\gamma_m)$ evaluated at $\dot\gamma_{m,0}$. Orange circles ({\color{Orange} $\bullet$}): estimate accounting for the non-zero $\epsilon$ using the measured value of $\epsilon=0.55$ in Eq.~3\label{eqn:tau_min}. Open red circles ({\color{Red} $\circ$}): estimate further accounting for the change in $\dot\gamma_m$ with $\epsilon$ and $B$ by evaluating Eq.~3\label{eqn:tau_min} at $\dot\gamma_m$ calculated from Eq.~\ref{eqn:dotgammamin}. Solid red circles ({\color{Red} $\bullet$}): Eq.~3\label{eqn:tau_min} at the measured $\dot\gamma_m$. Solid green circles ({\color{Green} $\bullet$}): prediction from Eq.~3\label{eqn:tau_min} at $\dot\gamma_m$ calculated from Eq.~\ref{eqn:dotgammamin} using only data obtained at $B=0$ and the lower shear thinning regime for larger $B$. Open green circles ({\color{Green} $\circ$}): evaluating Eq.~3\label{eqn:tau_min} at $\dot\gamma_m$ calculated from Eq.~\ref{eqn:dotgammaminfit} with $\alpha=1/2$. Dotted green line: fit of the open green circles indicating the phase boundary between shear thinning and Newtonian regimes. }

46 — 0908.1334

\caption{\label{fig:exp_conv}(Color online.) Convergence of the $1/\expect{\AO}$ expansion: ground-state energy difference $E^{(a/b)}_{i-1}-E^{(a/b)}_i$ as a function of the order $i$ for $\nuc{O}{18}$ \symbolcircle[FGBlue], $\nuc{Ni}{64}$\symbolbox[FGRed], and $\nuc{Sn}{120}$\symboldiamond[FGGreen]. }

\caption{\label{fig:SnXXX_D3}(Color online.) Theoretical three-point binding energy differences in the tin isotopic chain from $E^{(a/b)}$\,\symbolcircle[FGBlue] and $\widetilde{E}^{(b)}$\,\symbolbox[FGRed]\,, compared to experimental values\symbolcross \cite{Audi:2002rp}. }

\caption{\label{fig:SnXXX_DEhf}(Color online.) Ground-state energy differences for the tin isotopes in HF+EFA: $\widetilde{E}^{(b)}_{HF}-E^{(a)}_{HF}$ \symbolcircle[FGBlue] and $E^{(b)}_{HF}-E^{(a)}_{HF}$ \symbolbox[FGRed]. Calculations were done with 11 major shells. }

\caption{\label{fig:SnXXX_VAP}(Color online.) Top: Ground-state energy differences $\widetilde{E}^{(b)}_{PNP}-E^{(a)}_{PNP}$ of the tin isotopes from VAP calculations (note the scale in $\eV$). Bottom: Total pairing energies of the tin isotopes from VAP calculations with $E^{(a)}_{PNP}$ \symbolcircle[FGBlue] and $\widetilde{E}^{(b)}_{PNP}$ \symbolbox[FGRed]. All calculations were done with 11 major shells. }

47 — 0908.3260

\caption{ (Color online) (a): Generated turbulent state within the condensate edge (blue shading), %with a dimensionless vortex line density of $0.79$ at $t_0=1.9$ and $\gamma = 0$; the maximum density (used to determine the condensate edge) is $n=4.388 \times 10^{-3}$. (b) Evolution of normalized total length $L(t-t_0)/L_{0}$ for $\gamma = 0$ (black asterixes, {\color{black}{$\ast$}}) $0.015$ (purple pluses, {\color{AsPurple}{$+$}}), $0.03$ (blue circles, {\color{AsLightBlue}{$\circ$}}) and $0.06$ (red diamonds) corresponding to the same initial phase imprinting on a $128^3$ grid; %with the $\gamma=0$ exponential fit shown (solid line); the inset also shows the initial increase of $L$ for $t < t_0$ and the exponential fit for $\gamma=0$ (solid green line). }

\caption{ (Color online) Left: 3D log-linear velocity PDFs corresponding to Fig.\1 ($\gamma=0$): $\log_{10}[\text{PDF}(v_{i})]$ vs. velocity component $v_{i}$, for $v_{x}$ (blue circles, {\color{blue}{$\circ$}}), $v_{y}$ (red triangles, {\color{red}{$\bigtriangledown$}}) and $v_{z}$ (green asterixes, {\color{green}{$\ast$}}) , yielding following power law coefficients respectively $b = -3.27\pm0.04,\,\,-3.54\pm0.06,\,\,-3.57\pm0.06$. Corresponding $\log_{10}[\text{gPDF}(v_{i})]$ plots shown by black dotted ($\cdots$, $v_{x}$), dash-dotted (-$.$, $v_{y}$) and uniform (--, $v_{z}$) lines, which almost overlap. The velocity components are only sampled within the condensate edge (defined at 25\% of the maximum density) thus excluding the outer region where the density drops to zero. Corresponding statistical values: $\sigma^{2}_{i}=4.1,\,\,3.6,\,\,4.2$ and $\tilde{\mu}_{i}=\left(4.6,\,\,4.5,\,\,4.6\right)\times10^{-2}$. Right: Log-log plot of the same PDFs; the solid line is the power law fit to $v_x$. }

\caption{ Density (left) and velocity PDFs (right) for a 2D BEC at $t_0=4.4$ and $\gamma=0$. The maximum density is $n=2.057 \times 10^{-3}$. Left: $86$ vortices (core radius $\approx 2.66$) can be identified. Right: Plots of $\log_{10}[\text{PDF}(v_{i})]$ and the corresponding $\log_{10}[\text{gPDF}(v_{i})]$ vs $v_{i}$, where the $v_{x}$, $v_{y}$ velocity components (sampled within the condensate edge) are given by blue circles ({\color{blue}{$\circ$}}) and red triangles ({\color{red}{$\bigtriangledown$}}); corresponding gPDF's displayed by black dotted ($\cdots$) and dash-dotted (-$.$) lines. Statistical values: $\sigma^{2}_{x}=\sigma^{2}_{y}=2.0,\, \tilde{\mu}_{x}=5.8\times10^{-2}, \,\tilde{\mu}_{y}=1.3\times10^{-2}$. }

48 — 0908.4242

\caption{\textbf{Excitation spectra.} \textbf{a} Excitation energies with respect to momentum transfer $k_{\mbox{\scriptsize Bragg}}$ for harmonic trapping potential({\color{red} \ding{115}}) and lattice depths $V_0 = 3\,\mathrm{E_r}$({\color{green} \ding{108}}), $7\,\mathrm{E_r}$({\color{blue} \ding{110}}) and $11\,\mathrm{E_r}$({\color{magenta} \ding{117}}): The solid lines show calculations including interactions. Statistical errors always lie within the data points. \textbf{b} Excitation energies versus momentum transfer at higher lattice depths $V_0 = 7\,\mathrm{E_r}$ and $11\,\mathrm{E_r}$: The dashed lines show the single-particle bandstructure while the solid lines show the results of a Bogoliubov-approximated Bose-Hubbard calculation (see text). Error bars give the statistical uncertainty in energy, the uncertainty in momentum transfer lies within the data points. }

\caption{\textbf{Resonance spectra and corresponding bandstructure.} \textbf{a} Relative population in the central peak versus laser detuning $\delta$ at $k_{\mbox{\scriptsize Bragg}} = 0.93 \frac{{\pi}}{a}$ for lattice depths $V_0 = 3\,\mathrm{E_r}$({\color{green} \ding{108}}), $7\,\mathrm{E_r}$({\color{blue} \ding{110}}): The two main peaks of each spectrum correspond to the first and second band, respectively. Fit functions are given by a sinc function as the system's response to a square pulse. \textbf{b} Resonance position (first and second band) with respect to momentum transfer $k_{\mbox{\scriptsize Bragg}}$ along the [1,1]-direction: Dotted lines correspond to a single-particle bandstructure calculation. Solid lines show the excitation energies considering interactions based on a numerical solution of the Bogoliubov-de Gennes equation for the second band and following equation (\ref{burni}) for the first band. Dashed lines show a numerical solution of non-interacting particles with an attached external confinement (see text). Error bars give the statistical error in energy, the uncertainty in momentum transfer always lies within the data points. }

\caption{\textbf{Shift of resonance position with density.} Resonance position versus transversal lattice depth: The lattice depth in 2D is always $V_0 = 5\,\mathrm{E_r}$, and the measurements are performed at three momenta $k_{\mbox{\scriptsize Bragg}} = 0.64 \frac{{\pi}}{a}$({\color{blue} \ding{110}}), $k_{\mbox{\scriptsize Bragg}} = 0.71 \frac{{\pi}}{a}$({\color{green} \ding{108}}) and $k_{\mbox{\scriptsize Bragg}} = 0.93 \frac{{\pi}}{a}$({\color{red} \ding{115}}). The inset shows resonance frequency versus atom number for a 2D({\color{magenta} \ding{115}}) and a 3D lattice({\color{cyan} \ding{108}}) with transversal lattice depth $V_T = 5\,\mathrm{E_r}$ at $\Theta = 45^{\circ}$ and otherwise identical parameters. Error bars show the statistical uncertainty in energy.}

\caption{\textbf{Resonance position versus atom number with respect to the excitation fraction.} Excitation energy versus atom number at $k_{\mbox{\scriptsize Bragg}} = 0.71 \frac{{\pi}}{a}$ for harmonic trapping potential: Measurements are performed at excitation fractions of $75\%$({\color{red} \ding{115}}), $45\%$({\color{blue} \ding{110}}), $20\%$({\color{green} \ding{108}}), $6\%$({\color{magenta} \ding{117}}) and $3\%$({\color{black} \ding{115}}), respectively. The inset shows the corresponding data in a 2D lattice geometry at a lattice depth $V_0 = 7\,\mathrm{E_r}$. Measurements are performed at excitation fractions of $80\%$({\color{red} \ding{115}}), $45\%$({\color{blue} \ding{115}}), $23\%$({\color{green} \ding{108}}) and $8\%$({\color{magenta} \ding{117}}), respectively. Error bars show the statistical uncertainty in energy.}

49 — 0909.0082

\caption{(color online) Feedback cooling with varying feedback gain. {\bf A} and {\bf B}: In-loop and out-of-loop transduction spectra. %A Lorentzian fit to the Brownian motion allowed characterization of the mechanical oscillator, yielding a linewidth of $\Gamma=11.5$~kHz and mechanical quality factor of $Q_m = 545$ limited by gas damping in atmosphere. {\bf C}: Temperature inferred using in- and out-of-loop transduction signals. The red circles \textcolor{red}{$\medbullet$} and blue squares \textcolor{blue}{$\blacksquare$} respectively denote out- and in-loop temperature inferences; the solid red curve and dashed blue curve respectively denote the theoretical predictions of {\it actual} mechanical oscillator temperature and inferred in-loop temperature\cite{note}.}

50 — 0909.0790

\caption{Kinetics of the indirect exciton PL profile after the laser excitation onset. The measured (a) and calculated (b) cross-sections of the indirect exciton PL across the diameter of the inner exciton ring as a function of time. The measured (c) and calculated (d) indirect exciton PL intensity at the center of the laser excitation spot ({\color{red}{red $\blacktriangledown$}}) and at the inner ring radius $r = r_{\rm ring} \simeq 12\,\mu$m ({\color{blue}{blue $\medbullet$}}) where the PL maximum signal occurs, as a function of time. The time integration window $\delta t = 4$\,ns for each profile (a), (b) and each point (c), (d). The times$t=0$\,ns and$t=500$\,ns refer to the onset and termination of the rectangular laser excitation pulse. The laser excitation profile is shown by the thin dotted line in (a).}

\caption{Kinetics of the indirect exciton PL profile during the laser excitation termination. The measured (a) and calculated with Eqs.\,(\ref{diff})-(\ref{pl}) (b) spatial profiles of the PL signal from indirect excitons across the inner ring at the times $t = 500$\,ns (dash-dotted line), 504\,ns (dotted line), and 568\,ns (solid line). The times$t=0$\,ns and$t=500$\,ns refer to the onset and termination of the rectangular laser excitation pulse. The measured (c) and evaluated numerically (d) contrast of the PL-jump$(I_{\rm max} - I_{\rm laser\,on})/I_{\rm laser\,on}$ against the radial coordinate. The laser excitation profile is shown by the dotted line in (c). The measured (e) and calculated (f) PL intensity at the center of the laser excitation spot ({\color{red}{red $\blacktriangledown$}}) and at the radial distance where the PL maximum intensity occurs, $r = 12\,\mu$m, ({\color{blue}{blue $\medbullet$}}) as a function of time. Insets: The contrast of the PL-jump (d) and the PL-jump at the center of excitation (f), evaluated with Eqs.\,(\ref{diff})-(\ref{pl}) without time integration to match 4\,ns experimental resolution. Apart from the insets, each calculated curve is smoothed by the device resolution function with the time integration window$\delta t = 4$\,ns to match the experimental conditions.}

\caption{Modelling of the transient dynamics of the exciton temperature $T$ (a), density $n_{\rm x}$ (b), optical lifetime $\tau_{\rm opt}$ (c), and in-plane diffusion coefficient $D_{\rm x}$ (d) for the termination of the laser excitation. The radial coordinate is $r = 0$ (solid line), $r_{\rm ring}/2$ (dotted line), $r_{\rm ring}$ (dashed line), and $3r_{\rm ring}/2$ (dash-dotted line). Inset: The calculated spatial profile of the exciton temperature, $T = T(r)$, for $t = 500$\,ns (solid line) and$t = 504$\,ns (dotted line).}

51 — 0909.0805

\caption{{\bf Platonic solid measurement schemes.} Measurement axes ${\bf u}_k$ are defined by the Bloch-space directions through antipodal pairs of vertices of regular figures for $n=2,3,4,6$ and $10$ measurements. The figures are: (a) Square, $n=2$; and the four suitable Platonic solids: (b) Octahedron, $n=3$; (c) Cube, $n=4$; (d) Icosahedron, $n=6$; and (e) Dodecahedron, $n=10$. The {\red$\bullet$} symbols show the orientations of pure states in optimal cheating ensembles for two-qubit Werner states. In (d) and (e) these states align with the measurement axes ({\em vertices}), but in (a), (b) and (c), they have the {\em dual} arrangement, on the face-centres, similar to the situation in random access codes \cite{spekkensPRL}.}

52 — 0909.0909

\caption{Theoretical phase shift for the \blue\transition with zero magnetic field and a linearly polarized probe. It takes into account the three different$F' = 7/2,\ 9/2$ and $11/2$ levels of $^1\!P_1$, spanning over 60~MHz around their average frequency (center of the plot). The phase shift is represented for equally populated $m_F$ states (solid red curve) and spin-polarized atoms in $m_F = 9/2$ or $m_F = -9/2$ states (dashed blue curve). For a 90~MHz detuning, these phase shifts are comparable and amount to a few tens of mrad with a typical number of $N=10^4$ atoms.}

53 — 0909.1063

\caption{(Color online) Detector triggering and software efficiencies for the single (\textcolor{green}{\ding{115}}) and dual phase (\textcolor{blue}{\ding{108}}) runs. The \textcolor{red}{\ding{110}} points show the ideal case when the PMTs and the software have 100\% efficiency for a single photoelectron.}

\caption{(Color online) Scintillation efficiency for nuclear recoils relative to that of 122 keV gamma rays in LXe at zero field, comparing this work (\textcolor{DarkBlue}{$\medbullet$}) to previous measurements from Arneodo (\textcolor{black}{$\bigtriangleup$}) \cite{Arneodo:00}, Akimov (\textcolor{DarkGreen}{$\Box$}) \cite{Akimov:02}, Aprile (\textcolor{blue}{\ding{73}}) \cite{Aprile:05}, Chepel (\textcolor{red}{$\Diamond$}) \cite{Chepel:06} and Aprile (\textcolor{DarkRed}{$\bigcirc$})\cite{Aprile:08}. Also shown is the theoretical model (dashed line) explained in Section~\ref{sec:model}, which includes the Lindhard factor, an electronic quenching due to bi-excitonic collisions and the effect of escaping electrons.}

\caption{(Color online) Scintillation efficiency for nuclear recoils measured in this work (\textcolor{DarkBlue}{$\medbullet$}) and the theoretical model (dashed line) compared to the scintillation efficiency found from the neutron calibration data by the XENON10 (top shaded area) \cite{Sorensen:09} and the ZEPLIN-III (bottom shaded area) \cite{Lebedenko:08} collaborations.}

\caption{(Color online) Ionization yield as a function of recoil energy. Shown are the measured values in this work at 1.00~kV/cm (\textcolor{DarkBlue}{\ding{115}}) and 4.00~kV/cm (\textcolor{red}{$\medbullet$}), along with previously measured values at 0.10~kV/cm (\textcolor{black}{$\bigcirc$}), 0.27~kV/cm (\textcolor{red}{$\Box$}), 2.00~kV/cm (\textcolor{blue}{$\bigtriangleup$}) and 2.30~kV/cm(\textcolor{black}{$\Diamond$}) from \cite{Aprile:06}, error bars omitted for clarity. Also shown are the ionization yields calculated by comparing the XENON10 nuclear recoil data and the Monte Carlo simulations \cite{Sorensen:09} for single elastic recoils at 0.73~kV/cm, using two different methods (\textcolor{blue}{\ding{83}} and \textcolor{DarkGreen}{\ding{73}}). }

54 — 0909.2242

\caption{The partition $\lambda=(11,7,4,2,1,1,1,1,1,1)$ is an element of $B^H$ for $\asl_4$, since no box $b$ in~$\lambda$ is $A^H$-illegal. The color $\overline c(b)$ of each box $b$ is written inside $b$. We demonstrate the calculation of~$\tilde f_{\bar 2}(\lambda)$. The string of brackets $S_{\bar 2}^H(\lambda)$ has a ``('' for each {\color{green} ${\bar 2}$}-addable box and a ``)'' for each ${\color{green} {\bar 2}}$-removable box, ordered from left to right lexicographically, f\/irst by decreasing height$h(b)$, then by decreasing content~$c(b)$. The result is shown on the right of the diagram (rotate the page 90 degrees counter clockwise). Thus $\tilde f_{\bar 2}(\lambda)$ is the partition obtained by adding the box with coordinates $(x,y)= (10.5,0.5)$. The map $\Psi$ from Section~\ref{the-iso} takes $\lambda$ to $Y_{\bar 3, 11}^{} Y_{\bar 2, 8}^{} Y_{\bar 2,6}^{} Y_{\bar 3,5}^{} Y_{\bar 1, 5}^{} Y_{\bar 2, 10}^{} Y_{\bar 2, 12}^{-1} Y_{\bar 1, 9}^{-1} Y_{\bar 1, 7}^{-1} Y_{\bar 2, 6}^{-1} Y_{\bar 3, 11}^{-1}$. After reordering and simplifying, this becomes $Y_{\bar 2, 12}^{-1}Y_{\bar 2, 10}^{} Y_{\bar 1, 9}^{-1} Y_{\bar 2, 8}^{} Y_{\bar 1, 7}^{-1} Y_{\bar 1, 5}^{} Y_{\bar 3, 5}^{}.$ The string of brackets $S^M_{\bar 2}(\Psi(\lambda))$ is the same as the string of brackets $S^H_{\bar 2} (\lambda)$, except that a canceling pair $()$ has been removed. The condition that $\lambda$ has no $A^H$-illegal boxes implies that $S_{\bar \imath}^H(\lambda)$ and $S_{\bar \imath}^M(\Psi(\lambda))$ are always the same up to removing pairs of canceling brackets, which is essentially the proof that $\Psi$ is an isomorphism.\label{flat-crystal} }

55 — 0909.2688

\caption{ \textcolor{red}{ SNPs taken from the Table S4 of supplementary material of \cite{sladek} with tail area probability (obtained from permutation) smaller than $10^{-5}$, and if more than SNPs in a gene are significant at this level, only one SNP is chosen here. The first two columns list the gene/SNP name and the MAX3 value (based on the genotype counts given in the supplementary material of \cite{sladek}). (1) values of tail area probability provided by \cite{sladek}; (2) the best fit of $k$ when the values in column ``(1)" is used to fit a $\chi^2_k$ distribution; (3) estimation of the tail area probability of MAX3 by the distribution (for Max3) of 0.45+ 0.96$\chi^2_{k=1.7}$; (4) tail area probability of MAX3 by the exact enumeration of all possible combinations. } }

56 — 0909.4628

\caption{ \leg{Left} Ratio of negative displacements $\%\delta x^-$ in function of the reduced packing fraction $\varepsilon$, for three different forces : $F_1$ (\textcolor{bleu}{$\blacklozenge$}), $F_2$ (\textcolor{vert}{$\blacksquare$}) and $F_3$ (\textcolor{rouge}{$\blacktriangle$}). For clarity, only one point with a ratio of $0$ is shown. \inset Fluidisation force in function of the reduced packing fraction $\varepsilon = (\phi-\phi_J)/\phi_J$. Each point correspond to the moment where the ratio of negative displacements reaches $0$. These points are also marked with arrows in the left and middle figures. \leg{Middle} Ratio of negative displacements in function of the drag force $F$, for three different packing fractions : $\phi_1$ (\textcolor{rose}{$\triangle$}), $\phi_2$ (\textcolor{azur}{$\circ$}) and $\phi_3$ (\textcolor{orange}{$\square$}). \leg{Right} Applied drag force in function of the average velocity of the intruder in a $lin-lin$ plot at $\phi_1$. \inset Idem in a $lin-log$ plot for different packing fractions : $\phi_1$ (\textcolor{rose}{$\triangle$}), $\phi_2$ (\textcolor{azur}{$\circ$}) and $\phi_3$ (\textcolor{orange}{$\square$}). Fits are eye-guiding affine (dotted) or logarithmic (plain) behaviors. }

\caption{ \leg{Top and bottom-left} Average free volume fields around the intruder for three packing fractions: $\phi=0.8327$ ($\phi<\phi_J$, top-left), $\phi=0.8386$ ($\phi\simeq\phi_J$, top-right), $\phi=0.8405$ ($\phi>\phi_J$, bottom-left). \leg{Bottom-right} Average free volume $\langle \mathcal{V}_f \rangle$ before (\textcolor{rouge}{$\triangledown$}) and after (\textcolor{bleu}{$\vartriangle$}) the intruder in function of the reduced packing fraction. Error bars represent the standard deviation over time of the free volume in each point of space, averaged on the computation window. }

\caption{ Comparison of the evolution of dynamical lengthscales~: the cut-off of the distribution of the bursts of displacement's amplitudes $\xi$ is plotted for $3$ different forces (\textcolor{bleu}{$\Diamond$} $F_1$, \textcolor{vert}{$\Box$} $F_2$ and \textcolor{rouge}{$\triangle$} $F_3$) and compared to ($\bullet$) the long-range correlation length $\xi_4$, revealed from the dynamics of the grains without the intruder. The dotted and plain curves are guiding the eye like $\varepsilon^{-1}$. }

57 — 0909.5682

\caption{Positive monad bundles on the quintic. The entries marked in \textcolor{red}{red} are obstructions to a $\ZZZ$-action, see \autoref{sec:topology}. The last column $G_\text{3-gen}$ are free group actions such that the quotient has three generations.}

58 — 0909.5687

\caption{(Color online) Left: The crystallographic structure of \Litwo\comprises two AFM coupled CuO$_2$ spin-chains per unit cell running along the $b$-axis (orange {\large \color{orange} $\bullet$} -- Cu$^{2+}$, red {\large \color{red} $\bullet$} -- O$^{2-}$, bright blue {\large \color{brblue} $\bullet$} -- Li$^{+}$). The unit cell is indicated by the outer black cuboid. Right: the main intra- and inter-chain exchange paths, $J_1$, $J_2$, and $\tilde{J}_2$ marked by blue arcs and dashed lines, respectively. Notice the frustration introduced by an AFM inter-chain coupling for any non-FM in-chain ordering. }

59 — 0910.0866

\caption{Ratio $g$ between the number of cultural domains and the lattice area as function of the strength of the media influence for $L= 50 \left( \square \right)$ and $200 \left(\triangle \right)$. The solid line is the result of the extrapolation of the data to the limit $L^2 \to \infty$. The parameters are $F=5$ and $q=5$. \label{fig:1} }

60 — 0910.0940

\caption{\label{fig:cycle} Cycles of cooperation and defection in agent-based simulations. Top: dynamics in the $\nc$-$\cc$-plane for rewiring selection intensity $\alpha=30$. Bottom: degree distributions of cooperators (\full) and defectors (\dashed) for rewiring rate $p=0.8$ at the two turning points of the cycle. Simulation parameters as in \fref{fig:sim-oscill}.}

\caption{\label{fig:bif-p} Bifurcation diagram for the case of strong rewiring selection ($\alpha=30$). If rewiring is slow (small $p$), cooperators and defectors coexist in a stable steady state. At higher rewiring rates the stability is lost in a supercritical Hopf bifurcation (S), from which a stable limit cycle emerges. The limit cycle undergoes a homoclinic bifurcation in point H as it connects to a saddle point at $\nc=1$. The lines show the stable (\full) and unstable (\dashed) steady state, and the upper turning point of the limit cycle (red), computed in the low-dimensional model. Circles denote agent-based simulation results for $N=10^5$, $K=10^6$. The inset shows a blow-up of the bifurcation point S. See \fref{fig:sim-oscill} for parameters.}

\caption{\label{fig:bif-2par} Two-parameter bifurcation diagrams showing the dependence on the rewiring rate $p$ and the cost-to-benefit ratio \cite{Doebeli2005} $r=c/(2b-c)$ (left, for $\alpha = 30$), and the rewiring selection strength $\alpha$ (right, for $b=1$, $c=0.8$), resp. In region I, cooperation and defection coexist in a stationary state. A Hopf bifurcation line (\full) marks the transition to the oscillatory parameter region (II, shaded), which is bounded by a line of homoclinic bifurcations (\dashed) leading to asymptotic full cooperation (III). Stable full cooperation (IV) is reached via a transcritical bifurcation (\chain) if $r$ is low. All bifurcation lines meet in a codimension-2 Takens-Bogdanov bifurcation. See \fref{fig:bif-p} for additional parameters.}

61 — 0910.1073

\caption{Our results for the radiative lifetimes, $\tau_{\textrm{\footnotesize{eff}}}^{Exp}(T)$ of (a), the Rb $ns_{1/2}$ ($28 \le n \le 45$) states; (b), the $np_{3/2}$ ($34 \le n \le 44$) states; and (c), the $nd_{5/2}$ ($29 \le n \le 44$) states, at a nominal radiation temperature of $T=300$ K, plotted versus the effective principal quantum number, $n_{\textrm{\footnotesize{eff}}}$. The graphs also show theoretical predictions based on \eref{gammadef}, \eref{betbbr}, \eref{eq25}, and \eref{eq26}, using parameter values reported in a number of different papers. All the models use the BBR rates of Beterov \etal (equation (14) in \cite{bet09}); the solid red line (\full) uses values for $\tau_{\textrm{\footnotesize{s}}}$ and $\epsilon$ from Beterov \etal in \eref{eq25}; the purple dashed (\longbroken) \line uses Gounand's values for$\tau_{\textrm{\footnotesize{s}}}$ and $\epsilon$ \cite{goun79}; the blue dash-dot-dash (\chain) \line uses the values of He\etal for $a_0....a_3$ in \eref{eq26} \cite{he90}. The predictions using the values of He \etal for $\tau_{\textrm{\footnotesize{s}}}$ and $\epsilon$ in \eref{eq25} are essentially indistinguishable from those found using their values for $a_0....a_3$ in \eref{eq26}, except for the $nd_{5/2}$ curve, which is shown as a green dotted (\dotted) line.}

\caption{(a) Log-log plot of an extended set of rubidium Rydberg state zero-Kelvin decay rate measurements, $\Gamma_0^{Exp}$, comprising our results and those of \cite{goun80, mare80, goun76} for the $ns_{1/2}$ (full red diamonds, \fulldiamond), $np_{3/2}$ (open green circles, \opencircle), and $nd_{5/2}$ (blue crosses, $+$) states with $n \ge 10$. The range of $n_{\textrm{\footnotesize{eff}}}$ values is $6 \le n_{\textrm{\footnotesize{eff}}} \le 50$, and the range of $\Gamma_0^{Exp}$ values is $4 \times 10^3 \ \textrm{s}^{-1} \le \Gamma_0^{Exp} \le 4 \times 10^6 \ \textrm{s}^{-1}$. The previous results are the most precise values given in table V of \cite{theo84} that satisfy the criterion that $n \ge 10$. The zero-Kelvin decay rates are obtained from the data using \eref{tau0exp}, assuming that the calculations of Beterov and co-workers for the BBR depopulation rates are correct. The straight lines are fits of the data to \eref{eq25}, i.e., we fit the data to the equation $\Gamma_0^{Exp} = (1/\tau_{\textrm{\footnotesize{s}}}) \ n_{\textrm{\footnotesize{eff}}}^{-\epsilon}$. The full line (\full) is the fit to the $ns_{1/2}$ data, the dotted line (\dotted) to the $np_{3/2}$ data, and the dashed line (\dashed) to the $nd_{5/2}$ data. The dotted box shows our data, and is expanded in graph (b) below. (b) Inset of data presented in (a) between $24 \le n_{\textrm{\footnotesize{eff}}} \le 44$ showing only our results as a linear-linear plot. The $y$-axis is calibrated in units of 1000 s$^{-1}$. This plot also shows the fit lines to the entire data set shown in (a).}

62 — 0910.1797

\caption{ (color online) (a) An atom resolved STM image (46 x 46\AA, 2V, 0.2nA) showing two qubits created from pairs of DBs on a H-Si(100)2x1 surface, separated by 15.36\AA (Qubit 1) and 7.68\AA (Qubit 2). A schematic (left) shows the position of dangling bonds (red and green circles) on the Si surface. Black dashes represent silicon dimers. (b) A DB-DB$^-$ pair modeled as double-well potential, with the extra electron at the left well immediately after initialization to $\left|0\right\rangle$. (c) Relaxed ground state of the DB electrons after lattice relaxation has completed. } \label{fig:qubit} \end{figure} %\begin{figure} %\centering %\begin{tabular}{cc} %\epsfig{file=figure1.eps,width=0.5\linewidth,clip=} & %\epsfig{file=figure2.eps,width=0.5\linewidth,clip=} \\ %\epsfig{file=figure3.eps,width=0.5\linewidth,clip=} & %\epsfig{file=figure4.eps,width=0.5\linewidth,clip=} %\end{tabular} %\end{figure} The localized nature of the DB wavefunction and its energy level in the band gap allows us to formulate an electron-confinement model corresponding to a potential well accounting for the effect of the environment. %Such a potential well description must render the correct eigenstate energy %and orbital size, and must allow for electron excitation into the bulk conduction band of the crystal. Physically, localizing an extra electron at a DB invokes a lattice distortion due to interactions within the crystal yielding: (i) a 0.5eV upward shift of the DB$^-$ energy level relative to a neutral DB's to $\sim$0.85eV above the valence band edge (a change in the potential well resulting in weaker confinement and a lower ionization energy); and (ii) a local lattice deformation where the host Si atom at a DB$^-$ is raised by 0.3\AA~ from the plane of the surface. When the electron tunnels out of a DB$^-$, the lattice begins to relax. In Figs.~\ref{fig:qubit}(b,c) we depict a DB pair as an effective double-well potential with (b) an excess electron at the left well immediately after release from a biasing external field, as required for qubit initialization, and (c) after complete lattice equilibration when the potential landscape becomes symmetrical. Due to the localized extra charge, the double-well in case~(b) does not exhibit the symmetry of case~(c), and the DB energy is shifted upward at the left site. Consequently, during lattice relaxation, the coherent oscillation between the two DBs takes place between two wells of slightly different shapes, resulting in a periodic oscillation that is biased towards the `left' (excess electron spends more time on the left than on the `right'). Slow relaxation of the lattice will modify the electron oscillation and cause weak decoherence commensurate with the ratio of relaxation rate to oscillation rate. To study the effect of relaxation, we calculated tunneling rates for various separations by two different methods. For DB separations of $3.84$ and $7.72$\AA, tunnel splitting is determined to be $307.7$ and $88.7$meV, respectively, by time-dependent density-functional theory on cluster models~\cite{gaussian}. These correspond to tunneling rates of 9.3$\times$10$^{14}$Hz and 2.7$\times$10$^{14}$Hz, respectively. These points are marked by $\circ$ in the graph in Fig. \ref{fig:tunnelingdecoherence}. Tunneling rates for greater DB separations were calculated by the Wentzel-Kramers-Brillouin (WKB) method and are depicted by~$\square$. The black dashed line joining the calculated points shows an interpolation of the results obtained by the two methods. The red curve in Fig.~\ref{fig:tunnelingdecoherence} shows the calculated decoherence rate due to LA phonon relaxation of the charge-induced lattice distortion. Our calculation is based on a simple hydrogen-like model using the deformation potential approximation, previously used for charge qubits in bulk silicon~\cite{BM03}. We see that lattice relaxation via this mode occurs over several nanoseconds whereas the tunneling period for the DB-DB$^-$ pair with a few \AA~\!\! separation is close to 1fs. Other phonon modes both in bulk and at surface~\cite{TJS97} are less likely to couple to electron tunneling due to their discrete energy. Therefore, many coherent qubit oscillations will occur before lattice deformation is significant. We thus expect that Rabi type of oscillations will take place over many periods before decoherence sets in, and the advantage of closely spaced quantum dots is evident. \begin{figure}[tbp] \includegraphics[width=0.8\linewidth,clip=]{fig2_v5} \caption{ (color online) Tunneling rates of the excess electron in a charge qubit by time-dependent density-functional theory (black circles) and the WKB method (black squares) vs. DB separation $s$. % The dashed black line is an interpolation of the tunneling rates % calculated by the two methods. The red line depicts the calculated decoherence rate due to longitudinal-acoustical (LA) phonons. The shaded area indicates a region in which DBs are not tunnel coupled. } \label{fig:tunnelingdecoherence} \end{figure} \emph{Hamiltonian dynamics:---} As decoherence rates are orders of magnitude smaller than tunneling rates, the dynamics of DBs on the surface can be described by a Hamiltonian~$\hat{H}$ that acts upon the Hilbert space spanned by zero, one, or two electrons at each DB upon the silicon surface. On-site energy, electron tunneling (hopping), intra- and inter-DB Coulomb repulsion between electrons, and potential differences across the surface are all incorporated into~$\hat{H}$. We consider any number of DBs on the surface, with each DB at a site labeled~$i$. Let~$E_\text{os}$ be the on-site energy of an electron at any DB, which includes a constant surface chemical potential offset, and~$\eta_i$ be a site-dependent energy correction due to local field effects. The slowly changing lattice deformation due to the excess electron and the potential well deformation due to external biasing fields can be incorporated into this $\eta_i$ parameter. The coherent tunneling rate between sites~$i$ and~$j$ is~$T_{ij}$, which depends on the separation~$r_{ij}$ between the two DBs. %Electrons experience Coulomb repulsion from each other, which can be modified by electron spin. For two electrons at the same site,~$U_i$ denotes the energy cost of putting two electrons of opposite spin at the same site~$i$, including the energy contribution due to screening fields. The cost of putting one electron with spin~$\sigma\in\{\uparrow,\downarrow\}$ at site~$i$ and another electron of spin~$\sigma'$ at site~$j$ is denoted~$W_{i\sigma j\sigma'}$\cite{note:params}. Tunneling between DB sites can be controlled by modifying the surface potential. For example two sites~$i$ and~$j$ can have a time-dependent potential difference of~$V_{ij}(t)$. For $\hat{c}_{i,\sigma}$ ($\hat{c}^\dagger_{i,\sigma}$) the annihilation (creation) operator for an electron with spin~$\sigma$ at site~$i$ and $\hat{n}_{i,\sigma}=\hat{c}^\dagger_{i,\sigma}\hat{c}_{i,\sigma}$ the number operator for electrons of spin~$\sigma$ at site~$i$, the potential difference operator between sites~$i$ and~$j$ is $2\hat{V}\equiv\sum_{i<j,\sigma}V_{ij}(\hat{n}_{i,\sigma}-\hat{n}_{j,\sigma})$. We now have all the terms required to express the Hamiltonian as an extended Hubbard model~\cite{Hub78}: \begin{align} \label{eq:generalHam} \hat{H} =&\sum_{i,\sigma}(E_\text{os}+\eta_i)\hat{n}_{i,\sigma} -\sum_{\stackrel{i<j}{\sigma}}T_{ij}(\hat{c}^\dagger_{i,\sigma}\hat{c}_{j,\sigma} +\hat{c}^\dagger_{j,\sigma}\hat{c}_{i,\sigma}) \nonumber \\ &+\sum_i U_i\hat{n}_{i,\uparrow}\hat{n}_{i,\downarrow} +\sum_{i<j,\sigma,\sigma'} W_{i\sigma j\sigma'}\hat{n}_{i,\sigma}\hat{n}_{j,\sigma'}+\hat{V}. \end{align} Decoherence is neglected but could be incorporated by adding to this Hamiltonian coupling terms for various phonon modes such as the LA mode that is responsible for relaxation of the lattice deformation. Here Hamiltonian~(\ref{eq:generalHam}) suffices as the qubit dynamics is much faster than the decoherence processes, and we work in a regime where this Hamiltonian can be simplified to one of coupled qubits with standard descriptions of Markovian qubit decoherence~\cite{CL81}. \emph{Qubit dynamics:---} Hamiltonian~(\ref{eq:generalHam}) describes dynamics for quite general configurations of DBs on the silicon surface. For quantum computing, we need to generate entanglement by applying time-dependent gate potentials for specific qubit separation and relative orientation on the Si surface. A highly specific pattern of DBs, corresponding to grouping DBs into pairs of nearby DBs, and relatively large separations between pairs, greatly simplifies~(\ref{eq:generalHam}). %As spins factor out hence can be ignored, the state of each DB pair is For DBs on the silicon surface, electron spin is preserved so can be neglected; hence the `left' state~$\left|0\right\rangle$ and `right' state~$\left|1\right\rangle$ form a qubit basis~\cite{note:nonorthogonality} with conjugate basis corresponding to $\left|\pm\right\rangle=\left(\left|0\right\rangle\pm\left|1\right\rangle\right)/\sqrt{2}$. For $N$ DB pairs, the Hamiltonian can be conveniently rewritten in the qubit basis as a linear combination of quantum gates and tensor products thereof: $\openone$, $X=\left|0\right\rangle\bra{1}+\left|1\right\rangle\bra{0}$ and $Z=\left|0\right\rangle\bra{0}-\left|1\right\rangle\bra{1}$. Whereas~$i,j$ designates DB sites, $\imath,\jmath$ denotes DB \emph{pair} sites (equivalently charge qubits). The Hamiltonian is now expressed as an operator sum that acts on and between DB pairs: $\hat{H}_\text{q}(t) =\kappa\openone+ \sum_{\imath=1}^N\left[T\hat{X}_\imath +\frac{1}{2}\Delta V_\imath(t)\hat{Z}_\imath % +\frac{1}{2}\sum_{\jmath=1}^NW_{\imath\jmath}^{-}\hat{Z}_\imath\otimes\hat{Z}_\jmath\right]$. +\sum_{\jmath<\imath}W_{\imath\jmath}^{-}\hat{Z}_\imath\otimes\hat{Z}_\jmath\right]$. Intra-qubit separation is constant for all qubits with~$U_0$ and~$W_0$ on-site and inter-site Coulomb interaction within each DB qubit, %$W$ a constant incorporating inter-qubit Coulomb interactions, and $W_{\imath\jmath}$ the inter-qubit Coulomb repulsion $W_{\imath\jmath}^{\pm}=W_{\imath\jmath}^\text{s}\pm W_{\imath\jmath}^\text{c}$ and $W_{\imath\jmath}^\text{s}$ ($W_{\imath\jmath}^\text{c}$) is the inter-site Coulomb interaction between the same (cross) sites of two DB pairs $\imath$ and $\jmath$. Then $\kappa=N(3E_\text{os}+3\eta+U_0+2W_0)+\frac{9}{2}\sum_{\imath<\jmath}^NW^+_{\imath\jmath}$. Qubit-specific time-dependent potential-landscape tilting~$\Delta V_\imath(t)$ is incorporated into $\hat{H}_\text{q}(t)$, with~$T$ the intra-qubit electron tunneling rate. \emph{Initialization and readout:---} For our DB-DB$^-$ pair to be an effective charge qubit, initialization to a simple pure state and qubit-specific readout are critical; these are two of DiVincenzo's five criteria (the other criteria are scalability, a universal set of gates, long coherence times, with long coherence times established in the previous section)~\cite{DiV00}. Qubits are initialized in the $\left|0\right\rangle$ state by temporarily applying an electrostatic potential~$\Delta V_\imath(t)$ so that the left~DB is lower in energy thus attracting the pair's excess electron~\cite{BM03}. Steps toward this initialization procedure have been achieved experimentally for DB pairs~\cite{HBD+09}. When initialization is complete, the tilt of the electrostatic potential landscape is eliminated, and tunneling between the two DBs commences. Of course the lattice deformation due to excess charge in Fig.~\ref{fig:qubit}(c) is present during subsequent tunneling, but is expected to relax orders of magnitude slower than the tunneling rate hence gradually modify the tunneling dynamics with small decoherence over many cycles. Advances towards qubit-specific readout were achieved by STM detection of the excess charge~\cite{HBD+09}. Charge was observed to be preferentially localized at one site in a DB pair, which indicates that both state preparation on one side and readout of `left' vs `right' state is feasible. Both initialization and readout are slow processes. Initializing qubits can be slow so the preparation timescale is not a fundamental problem. As all measurements can be delayed until the end of the quantum computation at an acceptable cost of more qubits and gate operations, slow readout is a technical rather than fundamental obstacle. Of course fast measurements would be desirable, not only for error correction, but also to measure decoherence. %One approach to fast readout is to couple a nearby ancillary DB to the charge qubit %and tilt the potential landscape to the ancilla for the readout phase. %The goal would be to move the excess charge to the ancilla only if the %qubit is in the $\left|0\right\rangle$ and not otherwise. %Then a slow readout of the ancilla DB would determine the computational state, %$\left|0\right\rangle$ or $\left|1\right\rangle$. One approach to fast readout is to couple the charge qubit to a nearby molecule or single electron transistor (SET) and detect the changes in those when the qubit is in the $\left|0\right\rangle$ and not otherwise. The charging state of a DB was shown to affect the STM current through a nearby molecule implanted in the surface\cite{PLP+05}: the molecule's electronic structure can be Stark-shifted by the DB's excess electron. \emph{Universality:---} Hamiltonian~$H_\text{q}$ enables a universal set of gates~\cite{BBC+95}. Single- and two-qubit gates are effected by varying the inter-dot tunneling rate by tilting the potential landscape then rapidly turning off the tilting. Such fast and spatially precise control is beyond the capability of standard electronics but is conceivable by placing a suitable pattern of metallic nanowires near the surface and irradiating it with a laser pulse. The resulting electromagnetic field, created via plasmonic action~\cite{MA05}, can bias the surface with a temporal control comparable to the duration of the pulse (as short as femtoseconds). %with the brevity of the pulse bounded by the inverse of the laser carrier frequency: The laser carrier frequency should be low enough to avoid charging and discharging of DBs through excitation processes, thereby causing qubit losses. Different gates are effected via varying the bias duration by controlling the laser pulse. Scalability of our surface charge-qubit quantum computer follows the same arguments as for those cases, but of course better understanding of small-scale devices is required to assess scalability to large devices. Full universal quantum computing could proceed using a four-rail flying qubit model analogous to the one for electron-spin qubits in bulk silicon~\cite{HGFW06}. Alternatively the qubits can be stationary, and quantum computing could be implemented as a one-way quantum computer~\cite{RB01}. \emph{Summary:---} We show that closely-separated DB pairs on the silicon surface should behave as charge qubits, with coherence following from extreme miniaturization of qubits, truly reaching the atomic realm. Excellent coherence arises because the tunneling rate is extremely high due to atomic-scale proximity of DBs, whereas the major source of decoherence scales weakly with separation. The scaling advantage comes at the price of having to achieve rapid gating control. This fast control is a worthwhile challenge given the advantages of an all-silicon architecture for QC. Such fast gating control could be achieved by optical-plasmonic interfaces and precise positioning of metallic tips or nanowires. In the shorter term, experimental characterization of the decoherence for these charge qubits is of paramount importance. Whereas time-domain control is ultimately required, decoherence can be studied soon by fluorescence techniques: charge qubits are dipoles that will fluoresce in the terahertz regime, and decoherence can be learned from linewidths. In addition to weak decoherence, our scheme has another important advantage over other semiconductor charge qubit proposals: the charge qubits are on the surface rather than in the bulk medium, thus enabling more direct preparation, control, and readout. Furthermore steps toward initialization and readout of DB pairs have already been demonstrated so our scheme is leveraging off prior successes with DB quantum dot dynamics. We believe the findings outlined here could reinvigorate charge qubit prospects. \emph{Acknowledgments:---} This project has been supported by NSERC, MITACS, QuantumWorks, \emph{i}CORE, and the Southeast University Startup Fund. RAW is a CIFAR Fellow, and BCS is a CIFAR Associate. \begin{references} \bibitem{Gro97} L. Grover, \prl \textbf{79}, 325 (1997). % C. Zalka \textbf{60}, 2746 (1999). \bibitem{Sho94} P. W. Shor, Proc. 35th Annual Symp. on Found. of Comp. Sci. (Los Alamitos, CA: IEEE Computer Society Press, 1994), pp 124-34. \bibitem{Kan98} B. E. Kane, Nature (Lond.) \textbf{393}, 133 (1998). \bibitem{LD97} D. Loss and D. P. DiVincenzo, \pra \textbf{57}, 120 (1997); R. Vrijen et al., %E. Yablonovitch, K. Wang, H. W. Jiang, A. Balandin, V. Roychowdhury, % Tal Mor, and D. DiVincenzo \pra \textbf{62}, 012306 (2000). \bibitem{HGFW06} L. C. L. Hollenberg et al., \prb \textbf{74}, 045311 (2006); B. C. Sanders et al., New J. Phys. \textbf{10}, 125005 (2008). \bibitem{HFC+03} T. Hayashi et al., \prl. {\bf 91}, 226804 (2003). \bibitem{GHW05} J. Gorman, D. G. Hasko, and D. A. Williams, \prl{\bf 95}, 090502 (2005). \bibitem{NPT99} Y. Nakamura, Yu. A. Pashkin, and J. S. Tsai, Nature \textbf{398}, 786(1999); A. Wallraff et al., \prl \textbf{95}, 060501 (2005). \bibitem{HBD+09} M. B. Haider et al., \prl \textbf{102}, 046805 (2009). \bibitem{FHH04} T. Fujisawa, T. Hayashi, and Y. Hirayama, J. Vac. Sci. Technol. B \textbf{22}, 2035 (2004). \bibitem{Hub78} J. Hubbard, \prb \textbf{17}, 494 (1978). \bibitem{HGF+06} L. C. L. Hollenberg et al., \prb \textbf{74}, 045311 (2006). \bibitem{RB01} R. Raussendorf and H. J. Briegel, \prl \textbf{86}, 5188 (2001). \bibitem{ABW+07} S. E. S. Andresen et al., Nanolett. \textbf{7}, 2000 (2007). \bibitem{CL81} A. Caldeira and A. J. Leggett, \prl \textbf{46}, 211 (1981). \bibitem{note:nonorthogonality} In fact the wave functions corresponding to `left' and `right' occupation are not completely orthogonal, but the overlap is negligibly small for the case of two DBs separated by several \AA. \bibitem{BM03} S. D. Barrett and G. J. Milburn, \prb \textbf{68}, 155307 (2003). \bibitem{DiV00} D. P. DiVincenzo, Fortschr. Phys. \textbf{48}, 771 (2000). \bibitem{BBC+95} A. Barenco et al., \pra \textbf{52}, 3457 (1995). \bibitem{PLP+05} P.G. Piva et al., Nature \textbf{435}, 658 (2005). \bibitem{gaussian} M.J. Frisch, et al. Gaussian 03 Revision C.02 (Gaussian Inc., Wallingford, Connecticut, 2004) \bibitem{TJS97} H. M. Tutuncu, S. J. Jenkins, and G. P. Srivastava, \prb \textbf{56}, 4656 (1997). \bibitem{MA05} S.A. Maier and H.A. Atwater, J. Appl. Phys. \textbf{98}, 011101 (2005). \bibitem{note:params} Typical values of system parameters: $E_F$= 0.95eV, $E_{DB}$= 0.35eV, $E_{DB-}$= 0.85eV (all given with respect to the silicon valence band edge). Hubbard model parameters: $E_{os}$=-0.60eV, $U$=1.00eV; for a DB separation of 3.84\AA, $W$=0.58eV, $T$=0.307eV. \end{references} \end{document}}

63 — 0910.3153

\caption[]{(Color online) A snapshot of the simulation box. The system consists of 80 percent of particles of type A (blue) and 20\% of type B (red). Two atomistic walls (yellowish colors) confine the system along the $z$-direction (to the left and right of the image). Periodic boundary conditions are applied in the remaining $x$ and $y$ directions. The system size is $ L_x \times L_y \times L_z = 30 \times 30 \times 86$. It contains 92880 particles in total. \blue{The center of the coordinate system ($\vec{r}=\vec{0}$) is placed in the middle of the simulation box. Thus, $x\in [-L_x/2\;\; L_x/2]$, $y\in [-L_y/2\;\; L_y/2]$ and $z \in [-L_z/2\;\; L_z/2]$. $x$ is the flow direction, $y$ the neutral (vorticity) direction and $z$ the direction of the velocity gradient. All lengths are in LJ units.} }

64 — 0910.3650

\caption{Contributions to the ground-state energies in MBPT up to 30th order for different nuclei and model spaces: (a) \elem{He}{4} in a $12\hbar\Omega$ model space with $\hbar\Omega=20$ MeV (\symbolcircle[FGBlue]) and $32$ MeV (\symboldiamond[FGRed]); (b) \elem{O}{16} in $6\hbar\Omega$ with $\hbar\Omega=20$ MeV (\symbolcircle[FGBlue]) and $24$ MeV (\symboldiamond[FGRed]); (c) \elem{Ca}{40} in $4\hbar\Omega$ with $\hbar\Omega=20$ MeV (\symbolcircle[FGBlue]) and $24$ MeV (\symboldiamond[FGRed]). The upper panels depict the partial sum $E_{\text{sum}}(p)$ defined in Eq. \eqref{eq:runningsum} as function of the largest order $p$, the lower panels the modulus of the individual contributions $E^{(p)}$ on a logarithmic scale. The dashed horizontal lines in the upper panels indicate the NCSM ground-state energies for the respective nuclei and model spaces.}

\caption{Pad\'e approximants for the ground-state energies of (a)\elem{He}{4}, (b) \elem{O}{16}, and (c) \elem{Ca}{40} as function of the summed order $M+N$. The different symbols represent the diagonal approximants $E_{\text{Pad\'e}}(M/M)$ (\symbolcircle[FGBlue]), the super-diagonal approximants $E_{\text{Pad\'e}}(M-1/M)$ (\symboldiamond[FGRed]), and the sub-diagonal approximants $E_{\text{Pad\'e}}(M/M-1)$ (\symboltriangle[FGGreen]). The model space size $N_{\max}$ and the oscillator frequency $\hbar\Omega$ is quoted in the individual panels. The dashed horizontal lines indicate the NCSM ground-state energies for the respective nuclei and model spaces.}

65 — 0910.3673

\caption{\label{fig:PDPT15}Data from Fig.~\ref{fig:PD1513}(b) for $\delta = 0.15$ shown with variables pressure $P$ and temperature $T$. Crosses ({\color{blue} $\times$}) delimit the approximate region for a diffusion anomaly. Above and below that region, the diffusion coefficient decreases with pressure -- within the region the diffusion coefficient increases with pressure $P$. }

66 — 0910.4883

\caption{3C 454.3 Discrete Correlation Function (DCF) between the \gray and optical ($R$-band) magnitudes.}

\caption{Top panel: \gray 12-hr light curve above 100 MeV during the period between between 2007-12-05 and 2007-12-16 when the source was at 30 degrees in the \agile{} FoV. The vertical lines mark the time ($<3$ hr) of the exceptional optical event of 2007 December 12; bottom panel: $R$-band light curve obtained by WEBT (blue circles), REM (red circles), \MITSUME{} (green circles) during the period reported above.}

67 — 0910.5211

\caption{\label{fig3} (Color online) The universal ratio of double-to-single K-shell photoionization cross sections \cite{16,17}. Measurements by Huotari S {\it et al} \cite{8}: V, \textcolor{yellow}{$\newmoon$}; Cr, $\fullmoon$; Mn, \textcolor{blue}{$\blacktriangledown$}; Co, $\square$; Ni, \textcolor{green}{$\blacktriangle$}; Cu, \textcolor{red}{$\blacktriangledown$}; Zn, \textcolor{green}{$\bigstar$}. Measurements by Oura M {\it et al} \cite{25}: Ca, \textcolor{blue}{$\blacksquare$}; Ti, \textcolor{yellow}{$\blacktriangle$}; V, \textcolor{red}{$\newmoon$}.}

\caption{\label{fig4} (Color online) The universal ratio of double-to-single K-shell photoionization cross sections \cite{16,17}. Experimental data: Mg, \textcolor{red}{$\newmoon$}; Al, \textcolor{green}{$\blacktriangledown$}; Si, \textcolor{blue}{$\blacktriangle$} \cite{9,10}.}

68 — 0911.0296

\caption{\label{fig:pqmeos}\coloronl The normalized pressure (left panel) and the interaction measure (right panel) as a function of temperature. The model calculations (PQM model with various Polyakov-loop potentials and the QM model) are compared to lattice data ($N_\tau=8$, p4 and asqtad actions) from ~\cite{Bazavov:2009zn}. The solid lines correspond to larger pion and kaon masses as used in the lattice simulations while the dashed lines denote the results for physical masses.}

\caption{\label{fig:edensity}\coloronl The energy (left panel) and entropy density (right panel) similar to Fig.~\ref{fig:pqmeos}.}

69 — 0911.1291

\caption{Standard rheology of a 6~\% w/w carbon black suspension in a mineral oil. Viscoelastic moduli $G'$ ($\blacksquare$) and $G''$ ({\color{blue}$\bigcirc$}) (a) in a frequency sweep in the linear regime at $\sigma=0.5$~Pa, (b) in a stress sweep experiment at $f=1$~Hz and (c, d) at $\sigma=0.2$~Pa and $f=3$~Hz as a function of time after preshear at high speed during 60~s. The red dashed line in (b) indicates the yield stress $\sigma_y$, which is time and frequency independent.}

\caption{Large amplitude oscillatory shear experiments at $f=1$~Hz in a 6~\% w/w carbon black gel. (a) Temporal evolution of the deformation amplitude $\gamma_0$ for (from bottom to top) $\sigma=5$, 5.8, 6, 7, 8, 10, 13, 15, 18, 20, 23, 26~Pa. (b-d) Combined rheological and USV measurements for $\sigma=15$~Pa. (b) $\gamma_0$ and (c) $G'$ ({\color{black}$\bullet$}) and $G''$ ({\color{blue}$\bullet$}) as a function of time. (d) Spatiotemporal diagram of the ultrasonic speckle signal coded in linear grey levels. The USV sampling frequency is equal to the oscillation frequency $f=1$~Hz. Red dots indicate the boundary between solid-like and fluid-like regions. The red dashed line shows the time $\tau$ at which 90~\% of the gel across the gap is fluidized. (e) The fluidization time $\tau$ as a function of the stress amplitude $\sigma$. The black line is an exponential fit to the experimental data (only data above 8~Pa are considered in the fit). Region I: solid-like behaviour. Region II: solid-fluid coexistence. Region III: fluid-like response. Between the green and blue dashed lines, $\tau$ strongly deviates from the exponential behaviour.}

70 — 0911.1876

\caption{\label{fig_Walk} (a) Measurement of Fourier components $\left<\cos(kx)\right>$ and $\left<\sin(kx)\right>$ for a seven-step quantum walk. The data are obtained by varying the duration of the probe pulse for the ion prepared in the internal state $|+\rangle_z$ (left) or $|+\rangle_y$ (right) after completing the walk. The probability distribution is obtained by Fourier transforming a fit to the data (solid line). (b) Reconstruction of the symmetric part of the probability distribution $\langle\delta(\hat{x}-x)\rangle$ for up to 13 steps in the quantum walk. The blue dashed curve is a numerical calculation for the expected distribution within the Lamb-Dicke regime. The blue solid curve takes into account corrections to the Lamb-Dicke regime. In step 7, the dotted curve represents the full reconstruction using also the $\left<\sin(kx)\right>$ shown in (a).(c) Probability distribution of a five-step quantum walk after application of five additional steps which invert the walk and bring it back to the ground state.} \end{figure} Under the action of $H_{d}$, the ion's wave packet coherently splits in phase space along the $x$-axis. The two emerging wave packets $\psi_1^{(m)}$, $\psi_2^{(m)}$ are associated with the internal states $|\pm\rangle_x$. The length and the intensity of the pulse determine the width of the splitting. In our experiments we use a pulse of 40~$\mu$s with a Rabi frequency of $\Omega= (2\pi)\,68$~kHz to achieve a step size of $d = 2\Delta_x$. This step size makes the two resulting motional wave packets nearly orthogonal, $|\langle\psi_1^{(m)}|\psi_2^{(m)}\rangle|^2\approx 0.02$, but still allows for a large number of steps in phase space. Next, we perform a $\pi/2$-pulse acting on the carrier transition as a symmetric coin flip. This pulse creates an equal superposition of $\sigma_x$ eigenstates for both wave packets. These two pulses are repeated according to the number of steps to be carried out. \begin{figure} \includegraphics[width=7cm]{figure3.eps} \caption{\label{fig_Sigma}(Color online) (a) Width $w_x$ of the probability distribution in units of ground state size $\Delta_x$ as a function of the number of steps for a quantum (${\blacksquare}$) walk. The solid curve represents a full numerical simulation of the quantum walk as realized in the experiment. The width of the $x$-distribution for a classical random walk ($\textcolor[rgb]{1,0,0.00}{\bullet}$) increases more slowly and is described (solid red line) by eq.~(\ref{eqCl_sigma}). The data points (${\color{blue} \blacklozenge}$) show the measured width $w_p$ of the marginal distribution along the p-direction with $\Delta_p=\hbar/2\Delta_x$. (b) Average number of vibrational quanta after $N$ steps in the quantum walk measured by driving oscillations on the carrier transition. The solid line is based on a full simulation, the dashed line assumes the validity of the Lamb-Dicke approximation.} \end{figure} To measure the probability distribution along a line in phase space, we create two displaced copies of the state that are subsequently interfered. For this, we use of another state-dependent displacement operation $U_p=\exp(-ik\hat{x}\sigma_x/2)$~\cite{Wallentowitz:1995, Gerritsma:2010}. A measurement of $\sigma_z$ following the application of $U_p$ is equivalent to measuring the observable \beq\label{eqFourier} O(k)=U_p^\dagger\sigma_z U_p=\cos(k\hat{x})\sigma_z+\sin(k\hat{x})\sigma_y, \eeq with the usual position operator $\hat{x}=(a^\dagger + a) \Delta_x$ on the initial state. The propagator $U_p$ is obtained by setting $\phi_+$ and $\phi_-$ in $H_{D}$ to 0. Here, $k=2\eta\Omega_p t/\Delta_x$ is proportional to the interaction time $t$. If the ion's internal state is $|+\rangle_z$, we have $\langle O(k)\rangle=\langle\cos(k\hat{x})\rangle$ and for $|+\rangle_y$, we have $\langle O(k)\rangle=\langle\sin(k\hat{x})\rangle$. A Fourier transformation of these measurements yields the probability density $\langle\delta(\hat{x}-x)\rangle$ in position space which for a pure state $|\Psi\rangle$ amounts to $|\Psi(x)|^2$. Furthermore, we have that $\left.\frac{d^2}{dk^2}\langle O(k)\rangle\right\vert_{t=0}\propto\langle\hat{x}^2\sigma_z\rangle$~\cite{Lougovski:2006}. For eigenstates of $\sigma_z$, the initial curvature of the expectation value $\langle O(k)\rangle$ thus gives the width of the probability distribution $w_x$. The quantum walk entangles internal and motional degrees of freedom. Its analysis, however, requires the preparation of pure internal states like $|+\rangle_z$ or $|+\rangle_y$. Therefore, we recombine all internal state populations in $|-\rangle_z$ before the measurement. To this end, the population in $\lvert +\rangle_z$ is transferred to $\lvert -\rangle_z$ after transferring the population in $\lvert -\rangle_z$ to the auxiliary state $\lvert D_{\rm 5/2}, m = 5/2\rangle$. A laser pulse at 854~nm excites the population from $\lvert D_{\rm 5/2}, m = 5/2\rangle$ to $\lvert P_{\rm 3/2}, m = 3/2 \rangle$ from where it spontaneously decays to $\lvert -\rangle_z$. The efficiency of this pumping process is $>99$~\%, limited by a small branching ratio to the $D_{3/2}$-state. Only after the recombination step, we prepare the internal state required for measuring the even or odd Fourier components of (\ref{eqFourier}). Due to the small Lamb-Dicke parameter, the probability of changing the motional state of the ion during the pumping steps is small and hardly affects measurements of observables in position space at all. %, and could be eliminated completely by analyzing the motional state associated with the internal states $|+\rangle_z$ and $|-\rangle_z$ separately. In the experiment we set $\Omega_{p}=(2\pi)\,26$~kHz and measure $\langle \sigma_z \rangle$ for probe times between 0 and 300~$\mu$s in order to reconstruct the probability distribution $\langle\delta(\hat{x}-x)\rangle$ for different numbers of steps $N$. Since the walk is symmetric, it is in principle sufficient to measure only the even components of~(\ref{eqFourier}). For a seven-step walk, the measured odd and even Fourier components are displayed in Fig.~\ref{fig_Walk}(a). Panel (b) shows the reconstructed probability distribution $\langle\delta(\hat{x}-x)\rangle$ based on the even components for up to 13 steps. The uneven terms were checked to be close to zero for each number of steps $N$. The dashed lines in the plots are numerical simulations based on the Lamb-Dicke approximation. These lines deviate from the reconstructed distribution for $N>7$ due to higher order terms in $\eta$ that are not taken into account in eq.~(\ref{H_displace}). The solid lines are based on a numerical simulation using all orders. A similar difficulty occurs in the measurement of observables based on eq.~(\ref{eqFourier}). For this reason, the reconstruction is not accomplished by a direct Fourier transformation of the data. Instead, we apply a constrained least-square fit based on convex optimization~\cite{cvx:2009} capable of handling higher-order corrections (see the EPAPS document for more information on the reconstruction process). To get smoother distributions additional constraints were invoked by the reconstruction algorithm. A physical constraint is given by the maximal kinetic energy a one-dimensional wave packet can have. An estimate for the kinetic energy can be determined by measuring the momentum distribution in the same way as the position distribution. By changing $\phi_{\rm -} $ to $\pi/2$ in the probe pulse, the operator $\hat{x}$ appearing in (\ref{eqFourier}) is replaced by an operator $\propto\hat{p}$. These measurements (see Fig.~\ref{fig_Sigma}(a)) indicate that the momentum distribution is not seriously affected during the walk, as expected for a pure displacement along the $x$-axis. A striking difference between classical and quantum walks is the reversibility of the latter. In the experiment we reversed a quantum walk after five steps. This was done by switching the phase of the following five displacement and coin flipping pulses by $\pi$. In this way the quantum walk is exactly reversed and the ion returns to the ground state. The corresponding reconstructed probability distribution shown in Fig.~\ref{fig_Walk}(c) closely resembles the one of the initial state and demonstrates once more the coherence of the quantum walk. To further highlight the differences between quantum and classical walks we also realized a classical walk by randomizing the phase between each step (while keeping the coin flip-displacement operator pair coherent for each individual step). The phase for each step was generated by a random noise generator. This mimics a completely mixed ensemble of measurement outcomes that behaves classically. A good way of quantifying the difference between the quantum and classical walks is by measuring the average width of the probability distributions. For a classical walk with a step size $d=s\Delta_x$ we have % \beq\label{eqCl_sigma} w_x=\Delta_x\sqrt{\frac{2s^2N}{\pi}+1}, \eeq % where the second term takes into account the initial width $\Delta_x$ of the probability distribution. By contrast, for a quantum walk the width goes as $w_x \sim N$ for high $N$. To measure $w_x$ for the random walk, the curvature of $\langle \sigma_z \rangle$ at short probe-time was analyzed. Quadratic fitting gives direct access to the width $w_x$. For the quantum walk, $w_x$ was obtained from the measured probability distributions. In Fig.~\ref{fig_Sigma}(a) the results of these procedures can be seen seen for both a quantum and a classical walk. \begin{figure} \includegraphics[width=8.5cm]{figure4.eps} \caption{\label{TwoIonWalk} Reconstructed probability distribution $\langle\delta(\hat{x}-x)\rangle$ for a two-ion quantum walk with up to 5 steps with a step size of $4\Delta_x$.} \end{figure} To avoid problems in the measurement of the motional state due to leaving the Lamb-Dicke limit for large numbers of steps, we implemented a method suggested in~\cite{Xue:2009}. Outside the Lamb-Dicke regime the coupling strength $\Omega_{n,n}$ on the carrier depends on the phonon number $n$ as $\Omega_{n,n} = \Omega_0 L_n(\eta^2)$. Here, $L_n(\eta^2)$ is the $n$-th order Laguerre polynomial. The mean phonon number $\left<n\right>$ is determined by a constrained least-square fit of the carrier Rabi flops with the number state distribution as a fit parameter. In Fig.~\ref{fig_Sigma}(b), the resulting average vibrational quantum numbers are shown. As expected for the quantum walk, we observe a quadratic dependence $\langle n\rangle \propto N^2$ on the number of steps. Finally, we extend the quantum walk concept by adding a second ion to the system~\cite{Andraca:2005}. In the two ion quantum walk we make use of the center-of-mass mode. To account for the second ion, all Pauli matrices $\sigma_i$ in eq.~(\ref{eq:QuantumWalk}) are replaced by $\sigma_i^{(1)}+\sigma_i^{(2)}$. This changes the coin from two sided to four sided, with three possible operations. The "side" belonging to the state $|++\rangle_x$ ($|--\rangle_x$) corresponds to a step to the right (left) while the sides belonging to the states $|+-\rangle_x$ and $|-+\rangle_x$ correspond to no step at all. The ions are prepared in the state $|++\rangle_y$ with a $\pi/2$-pulse leading to a symmetric walk. For the two ion quantum walk all pulses are applied to both ions simultaneously. The probability distribution of the center of mass mode is obtained in the same way as for a single ion. The results for a walk of up to 5 steps are shown in Fig.~\ref{TwoIonWalk}. Again, the distribution deviates strongly from the classical version and shows a faster spreading. In summary, we have implemented a quantum walk using trapped ions. An experimental technique was implemented to determine the probability distribution along a line in phase space. This method might have further applications in quantum optics experiments or quantum simulations~\cite{Gerritsma:2010}. We have highlighted the difference between a classical and a quantum walk and demonstrated the reversibility of the latter. The current limitation in number of steps is given by instabilities in the trap frequency leading to decoherence and by the change in the coupling strength due to high phonon numbers. Quantum walks are of importance as a primitive for quantum computation~\cite{Childs:2009} and in finding search quantum algorithms that outperform their classical counterparts~\cite{Hillery:2009}. As such the experimental implementation of the quantum walk serves as an important benchmark and points the way to further experiments. For instance, the implementation of a quantum walk with two ions opens up the interesting possibility to introduce entanglement~\cite{Andraca:2005} and more advanced walks. \section*{Appendix: Reconstruction of the probability density} For the reconstruction of the probability density $p(x)=\langle\delta(\hat{x}-x)\rangle$ of the motional quantum state $\rho_m$, we determine the expectation value of the observable \beq\label{eqFourier} O(k)=U_p^\dagger\sigma_z U_p=\cos(k\hat{x})\sigma_z+\sin(k\hat{x})\sigma_y \eeq by applying the unitary $U_p=\exp(-ik\hat{x}\sigma_x/2)$ to the state $\rho=|\Psi\rangle\langle\Psi|\otimes\rho_m$ and measuring the operator $\sigma_z$. The right-hand-side of eq.~(\ref{eqFourier}) is obtained by using the equality $\exp(i\theta\sigma_x)=\cos\theta\, I+i\sin\theta\,\sigma_x$ and $\sigma_i\sigma_j=\epsilon_{ijk}\sigma_k$ for $i\neq j$. In this way, we determine $\langle\cos(k\hat{x})\rangle$ by choosing $|\Psi\rangle=|+\rangle_z$, and $\langle\sin(k\hat{x})\rangle$ by $|\Psi\rangle=|+\rangle_y$ \cite{Wallentowitz:1995}. In principle, a Fourier transformation of $f(k)=\langle\cos(k\hat{x})\rangle+i\langle\sin(k\hat{x})\rangle$ is sufficient for obtaining the density $\langle\delta(\hat{x}-x)\rangle$ in position space. However, with a finite number of experiments, the expectation values can only be determined for a discrete number of $k$-values and these measurements do not yield the exact expectation values but rather estimates of them. As a consequence, a reconstruction of the density based on Fourier transformation gives unphysical probability densities that are not non-negative everywhere. To overcome this problem, we reconstruct $p(x)$ by a constrained least-square optimization based on convex optimization~\cite{cvx:2009}. We discretize the position space by using a suitable set of points $x_i$ and search among the probability distributions with $p(x_i)\ge 0$ for all $x_i$ the distribution that minimizes \begin{eqnarray} \label{eqLeastSquare} S&=&\sum_k\left(\sum_ip(x_i)\cos(kx_i)-C_k\right)^2\nonumber\\ &&+\sum_k\left(\sum_ip(x_i)\sin(kx_i)-S_k\right)^2 \end{eqnarray} where $C_k$ and $S_k$ are the experimentally determined estimates of $\langle\cos(k\hat{x})\rangle$ and$\langle\sin(k\hat{x})\rangle$, respectively. In our reconstruction, we use another physical constraint based on a measurement of the kinetic energy $\langle\frac{\hat{p}^2}{2m}\rangle$ %$\langle{p}^2/(2m)\rangle$. in the following way. For a wave function $\psi=A(x)e^{i\phi(x)}$, a lower bound on the kinetic energy is given by \begin{eqnarray} \langle\frac{\hat{p}^2}{2m}\rangle&=&\frac{\hbar^2}{2m}\int_{-\infty}^\infty dx ((A(x)\phi^\prime(x))^2+A^\prime(x)^2)\nonumber\\ &\ge& \frac{\hbar^2}{2m}\int_{-\infty}^\infty dx A^\prime(x)^2 \end{eqnarray} where differentiation with respect to $x$ is indicated by primes. For $p(x)=|\psi(x)|^2$, we then have because of $A(x)=p(x)^{\frac{1}{2}}$ and $A^\prime(x)=\frac{1}{2}p^\prime(x)p(x)^{-\frac{1}{2}}$ that \beq \label{eq:KineticEnergyConstraint} \langle\frac{\hat{p}^2}{2m}\rangle\ge \frac{\hbar^2}{8m}\int_{-\infty}^\infty dx \frac{p^\prime(x)^2}{p(x)}. \eeq This constraint is also valid for mixed quantum states. In our optimization algorithm, eq.~(\ref{eq:KineticEnergyConstraint}) is a convex constraint that excludes distributions $p(x)$ having excessive energies. It requires a measurement of $\langle \hat{p}^2\rangle$ which is obtained by setting $\phi_-=\pi/2$ in the bichromatic Hamiltonian generating the unitary $U_p$ in eq.(\ref{eqFourier}) and calculating $d^2/dk^2\langle O(k) \rangle$. Adding the constraint works particularly well for the states produced by the quantum walk. These states ideally do not have any phase gradients $\phi^\prime(x)$ in which case inequality (\ref{eq:KineticEnergyConstraint}) turns into an equality. Moreover, the optimization algorithm can also handle to some extent problems related to the validity of the Lamb-Dicke approximation. In this approximation, the bichromatic laser-ion interaction, which is used for the operation $U_p$ in the reconstruction measurement, is described by the Hamiltonian \beq \label{H_displace} H_D = \hbar\eta\Omega\,\sigma_{x}\otimes (a + a^{\dagger}). \eeq This Hamiltonian is strictly valid only for $\eta\rightarrow 0$ because it is based on a Taylor expansion $e^{i\eta(a+a^\dagger)}=I+i\eta(a+a^\dagger)+{\cal O}(\eta^2)$ of the atom-light interactions that neglects terms in $\eta$ of order two or higher. If resonant terms up to third order are taken into account, the Hamiltonian becomes \beq \label{H_displace_thirdorder} H_D = \hbar\eta\Omega\,\sigma_{x}\otimes \left[(a + a^{\dagger})-\frac{\eta^2}{4}\left((a+a^\dagger)\hat{n}+\hat{n}(a+a^\dagger)+1\right)\right]. \eeq Since this Hamiltonian no longer commutes with $\hat{x}$, it cannot be used instead of eq.~(\ref{H_displace_thirdorder}) for the reconstruction procedure described above. On the other hand, an analysis based on eq.~(\ref{H_displace}) yields wrong results for quantum walks for large number of steps where the created states no longer fulfil the Lamb-Dicke criterion. Fortunately, for these states, their potential energy $\propto\langle(a+a^\dagger)^2\rangle$ is much larger than their kinetic energy $\propto\langle(i(a-a^\dagger))^2\rangle$ which provides the justification for replacing $\hat{n}$ by $\frac{1}{4}(a+a^\dagger)^2$ in eq.~(\ref{H_displace_thirdorder}). Using this assumption, the Hamiltonian \beq \label{H_displace_thirdorder_v2} H_D = \hbar\eta\Omega\,\sigma_{x}\otimes \left[(a + a^{\dagger})(1-\frac{\eta^2}{8}((a+a^\dagger)^2+1))\right]\eeq becomes again diagonal in the real space basis. The reconstructed probability densities $\langle\delta(\hat{x}-x)\rangle$ shown in the paper are based on this Hamiltonian. \begin{acknowledgments} We gratefully acknowledge support by the Austrian Science Fund (FWF), by the European Commission (Marie-Curie program), by the Institut f\"ur Quanteninformation GmbH. E.S. acknowledges support from UPV-EHU GIU07/40, Ministerio de Ciencia e Innovaci\'on FIS2009-12773-C02-01, EuroSQIP and SOLID European projects. This material is based upon work supported in part by IARPA.\end{acknowledgments} \begin{thebibliography}{0} \bibitem{Galton:1889} F.~Galton, {\it Natural inheritence} (Macmillan, 1889). \bibitem{Pearson:1905} K. Pearson, Nature {\bf 72}, 294 (1905). \bibitem{Rayleigh:1905} Rayleigh, Nature {\bf 72}, 318 (1905). \bibitem{Aharonov:1993} Y. Aharonov, L. Davidovich, and N. Zagury, Phys.~Rev.~A {\bf 48}, 1687 (1993). \bibitem{Kempe:2003} J. Kempe, Cont.~Phys. {\bf 44}, 307 (2003). \bibitem{Dur:2002a} W. D\"ur, R. Raussendorf, V. M. Kendon, and H.-J. Briegel, Phys.~Rev.~A{\bf 66}, 052319 (2002). \bibitem{Travaglione:2002} B. C. Travaglione and G. J. Milburn, Phys.~Rev.~A {\bf 65}, 032310 (2002). \bibitem{Sanders:2003} B. C. Sanders, S. D. Bartlett, B. Tregenna, and P. L. Knight, Phys.~Rev.~A {\bf 67}, 042305 (2003). \bibitem{Karski:2009} M. Karski {\it et al.}, % L. F\"orster, J.-M.Choi, A. Steffen, W. Alt, D. Meschede, and A. Widera, Science {\bf 325}, 174 (2009). \bibitem{Schmitz:2009a} H. Schmitz {\it et al.}, %R. Matjeschk, C. Schneider, J. Glueckert, M. Enderlein, T. Huber, and T. Schaetz, Phys.~Rev.~Lett. {\bf 103}, 090504 (2009). \bibitem{Schreiber:2009} A. Schreiber {\it et al.}, %K.-N. Cassemiro, V. Potocek, A. Gabris, P. Mosley, E. Andersson, I. Jex, and C. Silberhorn, Phys. Rev. Lett. {\bf 104}, 050502 (2010). \bibitem{Xue:2009} P. Xue, B. C. Sanders, and D. Leibfried, Phys.~Rev.~Lett. {\bf 103}, 183602 (2009). \bibitem{Kirchmair:2009} G. Kirchmair {\it et al.}, %J. Benhelm, F. Z\"ahringer, R. Gerritsma, C. F. Roos, and R. Blatt, New~J.~Phys. {\bf 11}, 023002 (2009). \bibitem{Wallentowitz:1995} S. Wallentowitz and W. Vogel, Phys.~Rev.~Lett. {\bf 75}, 2932 (1995). \bibitem{Gerritsma:2010} R. Gerritsma {\it et al.}, % G. Kirchmair, F. Z\"ahringer, E. Solano, R. Blatt, and C. F. Roos, Nature {\bf 463}, 68 (2010). \bibitem{Lougovski:2006} P. Lougovski, H. Walther, and E. Solano, Eur.~Phys.~J.~D {\bf 38}, 423 (2006). \bibitem{cvx:2009} M. Grant and S. Boyd, CVX: Matlab software for disciplined convex programming (2009), %\urlprefix \url{http://stanford.edu/~boyd/cvx}. \bibitem{Andraca:2005} S. E. Venegas-Andraca, J. L. Ball, K. Burnett, and S. Bose, New.~J.~Phys. {\bf 7}, 221 (2005). \bibitem{Childs:2009} A. M. Childs, Phys.~Rev.~Lett. {\bf 102}, 180501 (2009). \bibitem{Hillery:2009} M. Hillery, D. Reitzner, and V. Buzek, arXiv:0911.1102 (2009). \end{thebibliography} \end{document} % % ****** End of file template.aps ****** }

71 — 0911.2858

\caption{\protect\includegraphics{psiplot0}}

\caption{\protect\includegraphics{psiplot5}}

\caption{\protect\includegraphics{chi1plot}}

72 — 0911.3399

\caption{A PHENIX Run-5 \pp{} at \stwohundred{} dijet event. Charged tracks and photons are shown at the bottom by a Lego plot. The distribution of filter output values of the event is shown at the top as a contour plot. The maxima in the filter density are reconstructed as jet axes, shown as red lines at the positions on the contour and Lego plots.}

\caption{A PHENIX Run-5 \CuCu{} at \sNNtwohundred{} dijet event at $\approx 20\%$ centrality. Charged tracks and photons are shown at the bottom by a Lego plot. The distribution of filter output values of the event is shown at the top as a contour plot. The maxima in the filter density are reconstructed as jet axes, shown as red lines at the positions on the contour and Lego plots.}

\caption{Comparison of the jet energy scale between the $\sigma = 0.3$ Gaussian filter ($p_T^\mathrm{filt}$) and the $R = 0.4$ SISCone ($p_T^\mathrm{SISCone}$, with $0.5$ overlap threshold) \cite{Salam:2007xv} for \pp{} collisions at \stwohundred{} using \textsc{pythia} \cite{Sjostrand:2006za} and central \AuAu{} collisions at \sNNtwohundred{} using \textsc{pyquench} \cite{Lokhtin:2005px} at the event generator level (no detector effects). The dashed line indicates the position of $p_T^\mathrm{filt} = p_T^\mathrm{SISCone}$.}

\caption{The PHENIX jet $P(p_T^{pp} | p_T)$ transfer matrix for \stwohundred{} and $\sigma = 0.3$ Gaussian filter, derived from the \textsc{geant} simulation of $\approx 1.6\times 10^7$ \textsc{pythia} events. The $p_T^{pp} < p_T$ region is dominated by $n$, $K_L^0$ energy loss.}

\caption{PHENIX Run-5 \pp{} at \stwohundred{} invariant jet cross section spectrum as a function of $p_T$. The shaded box to the left indicates the overall normalization systematic uncertainty, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties.}

\caption{PHENIX Run-5 \pp{} at \stwohundred{} invariant jet cross section spectrum as a function of $p_T$, with comparison to \cite{Abelev:2006uq}, next-to-leading order calculation from \cite{Jager:2004jh}, and \textsc{pythia} assuming $K = 2.5$. The shaded box to the left indicates the overall normalization systematic uncertainty, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties.}

\caption{The $\sigma = 0.3$ Gaussian filter \pT{} resolution with $\Delta p_T/p_T \propto 1/p_T$ (constant energy smearing) fits for different centralities of \CuCu{} collisions at \sNNtwohundred{}, evaluated by embedding PHENIX Run-5 \pp{} events into PHENIX Run-5 \CuCu{} events. The constant energy smearing behavior is expected for an (asymptotically) linear jet reconstruction algorithm.}

\caption{Efficiency for the $g_{0.1}' > 52\,(\mathrm{GeV}/c)^2$ fake jet rejection that is suitable for central \AuAu{} collisions at \sNNtwohundred{} for \pp{} collisions at \stwohundred{} using \textsc{pythia} \cite{Sjostrand:2006za} and central \AuAu{} collisions at \sNNtwohundred{} using \textsc{pyquench} \cite{Lokhtin:2005px} at the event generator level (no detector effects), compared to the \textsc{pythia} efficiency for $g_{0.1}' > 17.8\,(\mathrm{GeV}/c)^2$ that is used for \CuCu{} collisions.}

\caption{Efficiency for the $g_{0.1}' > 17.8\,(\mathrm{GeV}/c)^2$ fake jet rejection for different centralities of \CuCu{} collisions at \sNNtwohundred{}, evaluated by embedding PHENIX Run-5 \pp{} events into PHENIX Run-5 \CuCu{} events}

\caption{Run-5 \CuCu{} at \sNNtwohundred{} azimuthal correlation for $0$--$20\%$ centrality dijets in yields and with different $g_{0.1}'$ fake jet rejection thresholds. The nominal fake jet rejection threshold used is $g_{0.1}' > 17.8\,(\mathrm{GeV}/c)^2$.}

\caption{PHENIX Run-5 \CuCu{} at \sNNtwohundred{} invariant jet yield as function of $p_T^{pp}$, with comparison to the $\langle T_{AB}\rangle$ scaled \pp{} cross section. The shaded box to the left indicates the centrality-dependent systematic uncertainty in the normalization, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties.}

\caption{PHENIX Run-5 \CuCu{} at \sNNtwohundred{} \RAA{} derived from unfolding. The shaded box to the left indicates the \pp{}--\CuCu{} systematic uncertainty in the jet energy scale, shaded boxes to the right shows centrality dependent systematic uncertainty between embedding and unfolding, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties.}

\caption{Comparison between the PHENIX Run-5 \CuCu{} at \sNNtwohundred{} \RAA{} derived from unfolding (filled symbols) and embedding (open symbols). The shaded box to the left indicates the \pp{}--\CuCu{} systematic uncertainty in the jet energy scale, shaded boxes to the right shows centrality dependent systematic uncertainty between embedding and unfolding, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties. Note that the flatness of \RAA{} makes a comparison across different energy scales possible.}

\caption{Comparison between the central PHENIX Run-5 \CuCu{} at \sNNtwohundred{} jet \RAA{} derived from unfolding and the $\pi^0$ \RAA{}. The shaded box to the left indicates the \pp{}--\CuCu{} systematic uncertainty in the jet energy scale, shaded boxes to the right shows centrality dependent systematic uncertainty between embedding and unfolding, shaded boxes associated with data points indicate point-to-point systematic uncertainties, and error bars indicate statistical uncertainties. Note that while the flatness of \RAA{} makes a comparison across different energy scales possible, $\pi^0$ with $\langle z\rangle = 0.7$ has a different energy scale.}

\caption{PHENIX Run-5 \CuCu{} at \sNNtwohundred{} azimuthal jet-jet correlation with Gaussian fits for jets with $7.5\,\mathrm{GeV}/c < p_T^\mathrm{CuCu} < 11.5\,\mathrm{GeV}/c$}

\caption{Widths of Gaussian fit to the PHENIX Run-5 \CuCu{} at \sNNtwohundred{} azimuthal angular correlation for jets with $7.5\,\mathrm{GeV}/c < p_T^\mathrm{CuCu} < 11.5\,\mathrm{GeV}/c$}

73 — 0911.3523

\caption{(Color online). Effect of ions (a) Theoretical spectra for $\vert r \rangle~=~$48S$_{1/2}$ for ground state density~2.2~$\times~10^{10}$~cm$^{-3}$ for 0,~2 and 5~\% ion density showing a Stark shift in the two-photon resonance before significant suppression of the resonance. (b) Measured two-photon resonance for experimental scan (\textcolor{red}{$\blacksquare$}) shows suppression not due to ions. Reversing the probe scan direction to first create ions shows a visible Stark shift in the second part of the scan (\textcolor{blue}{$\bullet$}), showing the sensitivity to ions. Each point is the average of 20 measurements.}

74 — 0911.4057

\caption{Calculated OBSM phonon anomalies in La$_{2}$CuO$_{4}$: RIM (\dashddot), RIM with anisotropic DF's (\chain), 11BM (\dotted), 12BM (\broken) and $\Pi$-model (\full). The squares (\fullsquare) represent experimental results \cite{Pint98}.}

\caption{Calculated phonon dispersion of HgBa$_{2}$CuO$_{4}$ for model $\Pi$-3dim in the main symmetry directions $\Delta\sim(1,0,0)$, $\Sigma\sim(1,1,0)$ and $\Lambda\sim(0,0,1)$. The symbols $\circ$ and $\Box$ represent experimental results from \cite{d'Ast03} and \cite{Uchi04}, respectively. The circle $\fullcircle$ marks two breathing modes at $M$ and the half-breathing mode at $X$. The arrangement of the panels from left to right according to the different irreducible representations is as follows:$|\Delta_{1}|\Delta_{2}$ (\dotted), $\Delta_{4}$ (\full)$|\Delta_{3}|\Sigma_{1}|\Sigma_{2}$ (\dotted), $\Sigma_{4}$ (\full) $|\Sigma_{3}|\Lambda_{1}$ (\full), $\Lambda_{2}$ (\dotted)$|\Lambda_{3}|$. Imaginary frequencies are represented as negative numbers.}

\caption{Reconstructed experimental x-ray spectra $N(E)$ of (a) HgBa$_{2}$CuO$_{4}$ for $\vc{q}=\frac{\pi}{a}(\zeta,0,0)$ \cite{Uchi04} and (b) Bi$_{2}$Sr$_{2}$CuO$_{6}$ for $\vc{q}=\frac{2\pi}{a}(\zeta,0,0)$ \cite{Graf08} in arbitrary units, respectively. The experimental spectra for the different $\zeta$ are represented by the open and closed circles (\opencircle,\fullcircle). The according interpretations of the phonon frequencies fitted by a Voigt fit \cite{Uchi04,Graf08} are represented by open triangles (\opentriangle). Our calculated phonon frequencies are represented by closed triangles (\fulltriangle).}

\caption{Comparison of model $\Pi$-2dim (\dashed) and model $\Pi$-3dim (\full). The three leftmost panels belong to HgBa$_{2}$CuO$_{4}$ and the five rightmost panels belong to Bi$_{2}$Sr$_{2}$CuO$_{6}$. Only irreducible representations with a different phonon dispersion in model $\Pi$-2dim and model $\Pi$-3dim are shown. The arrangement of the panels from left to right according to the different irreducible representations is as follows: $|\Delta_{1}|\Sigma_{1}|\Lambda_{1}||\Delta_{1}| \Delta_{3}|\Sigma_{1}|\Sigma_{3}|\Lambda_{1}|$.}

\caption{Calculated phonon density of states (DOS) for (a) Hg$_{2}$BaCuO$_{4}$ and (b) Bi$_{2}$Sr$_{2}$CuO$_{6}$ for model $\Pi$-3dim (\full). For comparison, the DOS of the RIM is shown as dotted line (\dotted). The symbols $\times$ and $\bullet$ in (a) show the experimental data \cite{Renker96} at 30 and 300K, respectively, the symbols $\bullet$ in (b) represent the experimental data \cite{Parshin96} at 300K. All experimental data are in arbitrary units and shifted for better viewing.}

75 — 0911.4231

\caption{{\color{blue}} Experimentally recorded images of light scattered from two CO nematicons at various input powers $P$ in each beam.}

\caption{{\color{blue} (Video 1)} Experimental results of the temporal evolution of two CO solitons shown in Fig.~\ref{fig2} for power 13.5~mW.}

\caption{{\color{blue} (Video 2)} Experimental results for the temporal evolution of two CP nematicons shown in Fig.~\ref{fig3} for power 13.5~mW.}

76 — 0912.0472

\caption{\co Times scales as a function of packing fraction. Left: Distributions of $\tau$, the lag time between adjacent clusters; \textit{Main plot} $P(\tau>\tau_1)$ -- black dotted lines are exponential fits at large $\tau_1$. \textit{Inset} Pdf$(\tau)$ for the population of short lag times (see text for more details) -- black lines are indicative exponential decays. Right: $\tau_{1/2}$~(\textcolor{red}{\textbullet}) , $\tau_{cl}$(\textcolor{blue}{$\blacktriangleright$}), $\tau_S$~(\textcolor{green}{$\blacktriangledown$}), and $\tau_f$~({\tiny$\blacksquare$}). }

\caption{\co Length scales as functions of the packing fraction. Left: Cumulated Pdf of the clusters' number of particles $n_c$ for the $6$ packing fractions $\phi=0.758$ (red) to $0.802$ (blue); \textit{Inset} Mean value of $n_c$ over all clusters, as a function of $\phi$. Right: $\xi_{cl}$(\textcolor{blue}{$\blacktriangleright$}), $\xi_{1/2}$~(\textcolor{red}{\textbullet}), and $\xi_{ava}$~({\tiny$\blacksquare$}). Plain lines are guides for the eyes, dashed lines are extrapolations. }

77 — 0912.1868

\caption{Differential energy spectrum for the \gray{s} produced in the Earth's atmosphere taken in different $\Theta_{\mathrm{nadir}}$-bands. The spectrum of the limb (open circles) shows a power-law behavior. Also shown are \gray{} spectra for the inner parts of the Earth's disk (filled circles and points). Indicated are both statistical and systematic errors (grey band), the latter derived from studies of the Vela pulsar. The inset shows a comparison to previous data by the \sas\satellite~\citep{SASEarth}. To allow for a direction comparison, the limb-spectrum is also shown for a larger integration region ($58^{\circ} < \Theta_{\mathrm{nadir}} < 78^{\circ} $) corresponding to the \sas\measurement. The bottom panel shows the limb-spectrum multiplied by energy to the power 2.75.\label{fig::Spectrum}}

\caption{Comparison between the integrated column density with the LAT-detected gamma-ray intensity as a function of nadir angle for events above 3.6~GeV. Two different model atmospheric profiles have been used: a simple barometric atmosphere with scale height of 6.8~km~\citep{GaisserBook} shown as solid line and a more realistic atmospheric model from~\citep{NRLMSISE}. Shown are the measured gamma-ray intensity (solid circles) and the gamma-ray intensity deconvolved by the PSF of the median energy (4.7~GeV) of this data set (open circles). A good correspondence between the \gray{} intensity and the column density can be seen at angles $\Theta_{\mathrm{Nadir}} \gtrsim 68.3^{\circ}$. For these angles the atmosphere is thin enough (column density $< 10$ g cm$^{-2}$) so that no significant attenuation occurs and the \gray{s} are directly related to the amount of target material for the incoming cosmic rays. The absolute level of the \gray{} intensity has been scaled to match the column density for the barometric atmosphere model. \label{fig::atmosphere}}

\caption{A comparison of the measured gamma-ray intensity for the angular interval $68.6^{\circ} < \Theta_{\mathrm{nadir}}< 69.6^{\circ}$ (thin target case, gamma-ray attenuation is insignificant) with the scaled cosmic ray proton intensity (see text for details). The inset shows the ratio between the \gray{} data and the scaled proton data and shows the general good match between the two measurements. The data have been taken from~\citet{BESS} and \citet{ryan1972}. \label{fig::GammaProtonRatio}}

\caption{Energy intensity for the Earth-originating \gray{s} from keV to 500 GeV energies. The measurements from different instruments are not readily comparable since they have been taken in different points in the solar cycle and are also integrated over different areas of the Earth. Data points are taken from~\citet{Ajello2008, Frontera2007, Schwartz, Churazov, Gehrels1985, ryan1977, Share2001}. \label{fig::SED}}

78 — 0912.2216

\caption{Model Parameters for the 2007 \gray Flares of \source.}

\caption{ Bottom panel: SED relative to the 2007 September 11 data. Curves labeled "c I" and "c II" (gray and blue solid lines) show the two separate components. Data in light gray represent a previous low state plus the EGRET observation in \gray (see Lin et al. 1995). Top panel: two-components SED relative to the 2007 October flare (solid line), and after 1 day by radiative cooling (dotted line). In both panels black stars are simultaneous data (see Villata et al. 2008; Giommi et al. 2008; Chen et al. 2008), the black lines are the two-components models and the dashed red lines represent the best one-component models.}

\caption{Variability plane with the trajectories of 0716+714 based on simultaneous optical and \gray observations. Open circles are for 2007 September, and filled triangles for 2007 October data. The thick solid line represents the quadratic trend, expected from electrons injected or accelerating in the SSC framework; for comparison, the linear behavior expected from EC is represented by the thin dotted line.}

79 — 0912.2468

\caption{(Color online) A simple driven system, amenable to numerical calculations. (a) Energy as a function of position. A single particle occupies a periodic, one-dimensional energy landscape. The position coordinate is discretized into $N_x=32$ uniformly spaced positions per period. Energy is also discretized, $E(x) = \lfloor N_e (1+ \sin (2\pi x/N_x))/2 \rfloor/N_e$ where $x$ is position and $N_e=64$ is the number of discrete energy bins. Results do not change appreciably with a finer discretization in either position or energy. (b) The system is initially in equilibrium with an external heat bath~(\textcolor{blue}{$+$}). At each discrete time step, the particle attempts to move one step left, one step right, or remain in the same location with equal probabilities, and the move is accepted according to the Metropolis criterion~\cite{Metropolis1953}. Every $1/v$ time steps, the energy surface shifts one position to the right. To ensure fully time-reversible dynamics, we simulate $1/2v$ time steps, shift the potential, and simulate another $1/2v$ time steps before examining the non-equilibrium properties of the system. All figures are drawn in the rest frame of the potential. Eventually the spatial distribution across a single periodic image converges to a non-equilibrium steady state~(\textcolor{red}{$\times$}), approximated by Eq.~(\ref{eq:Pneq})~({\scriptsize \color{MyPurple}$\Box$}). (c) The excess mean work~({\scriptsize $\blacksquare$}) varies markedly as a function of starting position, whereas the excess work variance~(\textcolor{green}{$\circ$}) is relatively constant, and the sum of second- and higher-order cumulants $K_{x,\R}$~(\textcolor{MyOrange}{$\bullet$}) [Eq.~(\ref{eq:Kdef})] is even moreso, implying compensating behavior. Displayed results are for $\beta = 4$ and $v=1/4$. The reciprocal temperature $\beta$ is reported in inverse units of the energy difference between top and bottom of the potential.}

\caption{(Color online) Equilibrium~(\textcolor{blue}{$+$}), steady state~(\textcolor{red}{$\times$}) and approximate steady state~({\scriptsize \color{MyPurple} $\Box$}) [Eq.~(\ref{eq:Pneq})] probability distributions, for the system described in Fig.~\ref{fig:systemSine}, at various driving rates and temperatures. The quality of our approximate distributions, including overall normalization, deteriorates at low temperature and high driving velocity. The dotted box highlights the conditions shown in Fig.~\ref{fig:systemSine}. }

80 — 0912.2739

\caption{\label{fig:Re400fields} (color online). Localized traveling wave $\butw$ (a,b) and equilibrium $\bueq$ (c,d) solutions of plane Couette flow at $\Reynolds=400$, from \cite{Schneider2009}. The velocity fields are shown in the $y=0$ midplane in (a,c), with arrows indicating in-plane velocity and the \colorswitch{color}{gray} scale indicating streamwise velocity $u$: \colorswitch{blue/green/red}{black/grey/white} correspond to $u=-1,0,+1$. The $x$-averaged streamwise velocity is shown in (b,d), with $y$ expanded by a factor of three. }

81 — 0912.3000

\caption{Phase boundary for $a=0.5$ and $\eta\propto\sqrt{\rho}$. The critical points found via the Dickman and Dornic \textit{ et al.} algorithms are shown as black crosses and red diamonds, respectively. Both methods agree up to an accuracy of about 1\%. \red{The solid maroon line correspond to the theoretical prediction (see below) for the spinodal line whereas the blue squares correspond to its numerical counterpart (see below).} }

\caption{Plots of the average order parameter $\langle m \rangle$ as a function of $v$, for $\Gamma_0=0.10$ and for three different system sizes (see legend). The presence of a crossing \red{of the order parameter curves} indicates that the transition is discontinuous. Inset: probability distribution of the order parameter, for $v=v_c\simeq 1.21$ and $L=100$. }

\caption{Plots of the average order parameter $\langle m \rangle$ as a function of $v$, for $\Gamma_0=0.50$ and for three different system sizes (see legend). The presence of a crossing \red{of the order parameter curves} indicates that the transition is discontinuous. Inset: probability distribution of the order parameter, for $v=v_c\simeq 0.58$ and $L=100$.}

\caption{Contour lines in the $(v,\Gamma_0)$ plane for which $\tau$ equals, from left to right, 500, 100, 50 and 30. In order to plot the contour lines of the 0D model, we use $\beta=2.3\, v$ \red{(see below). The analytical prediction for the spinodal line is also plotted, and closely matches the curves for which $\tau=500$.}}

82 — 0912.3560

\caption{% (Color online) {\bf a} Optimal spatial configuration of $N=7$ sites offering fast, robust, and complete transport from input to output. {\bf b} Time evolution of the on-site probabilites $|\ovl{i}{\psi(t)}|^2$ generated by the Hamiltonian defined by {\bf a}. $i$ is either the input site \protect\includegraphics{article_pin_reresub}, the output node \protect\includegraphics{article_pout_reresub}, or an intermediate site \protect\includegraphics{article_p1_reresub}. At time $\stime'$, only the output site is populated. At intermediate times $t<\stime'$, the excitation is spread over several sites, leading to high values of the bipartite \protect\includegraphics{article_c2_reresub} and quadripartite \protect\includegraphics{article_c4_reresub} entanglement, see text. }

83 — 0912.4104

\caption{\label{fig2}(Color online) Schematic evolution of DPs in MBZ with increasing staggered flux $\delta$ under SMF-I. ``{\color{red}$\bullet$}'' and ``$\bullet$'' represent the original DPs in pristine graphene and the induced (additional) DPs respectively, while the arrows represent their moving directions. The black circles denote isotropic Dirac cones whereas the green (grey) ellipses denote anisotropic Dirac cones. The coordinates of the four corners of MBZ are ($\pm\pi/3, \pm\pi/\sqrt{3}$).}

84 — 0912.4687

\caption[Comparison of K-band Spectra]{\label{Kspec} Keck~II/OSIRIS $K$-band spectra of \sciencebina\and B, compared to IRTF/SpeX spectra of a low-gravity L3 dwarf (G196-3B, green), a field L3 dwarf (2MASS~1506+1321, magenta), a field L5 dwarf (2MASS~1507-1627, cyan) and an L4.5 dwarf with a dusty photosphere (2MASS~2224-0158, red). The spectra of\sciencebina\and B have been smoothed to the SpeX spectral resolution (R$\sim$2000). The spectra of field and dusty dwarfs are from \citet{cushing05}, and the G196-3B spectrum is from \citet{allers07}.}

\caption[NIR colors]{\label{colors} MKO $J-K$ (top) and $K-$\Lp\(bottom) colors as a function of spectral type for L dwarfs. Field dwarf $J-K$ colors \citep{knapp04} and $K-$\Lp\colors\citep{golimowski04} are shown as gray circles. The black squares show the colors of \sciencebina\and B (Table\ref{binphot}). The green diamond shows the color of G196--3B, a $\sim$100~Myr L3 dwarf \citep{cruz09}, synthesized from its near-IR spectrum. The open red triangles show the colors of 2MASS~~2224$-$0158 \citep{knapp04,golimowski04}, an L4.5 field dwarf with a dusty photosphere \citep{kirkpatrick00, cushing05}. \sciencebina\and B have red colors compared to normal field dwarfs of the same spectral type, which could be due to a low surface gravity (youth) or a dusty photosphere.}

\caption[Mass determination]{\label{hr} Plot of age vs. effective temperature (top) and log($L_{\rm{bol}}/L_\odot$) (bottom) for \sciencebina\and B with model iso-mass contours from\citet{burrows97} overlaid. The shaded regions show the range of possible ages for \sciencebin\from our strict age limits (12-790~Myr; light gray shading) and the age of G196--3B (20--300~Myr; gray shading).}

\caption[Mass Ratio determination]{\label{massratio} Plot of mass vs. log($L_{\rm{bol}}/L_\odot$) for model isochrones of 20, 75, 140, and 300~Myr \citep{burrows97}. The solid horizontal lines show our measured luminosities of \sciencebina\(red) and B (blue), while the shaded region represents the range of 1$\sigma$ uncertainty. While the absolute luminosities of \sciencebina\and B are quite uncertain, the relative luminosity (represented by the separation between the red and blue lines; 0.36$\pm$ 0.02~dex) is well determined. The bump in the isochrones at $\sim$0.013~$M_{\odot}$ is due to deuterium burning and means that for the range of possible ages and luminosities of \sciencebin, its mass ratio cannot be robustly determined. In fact, \sciencebin\could have a flux reversal due to deuterium burning, where the lower mass object is currently burning deuterium and is thus the more luminous component.}

\caption[Galactic Space Motion]{\label{uvw} $UVW$ space for \sciencebin\and known young moving groups, plotted using a right-handed coordinate system ($U$ positive toward the Galactic center). $UVW$ for \sciencebin\(solid black line) is calculated from its position, distance, and proper motion for a range of radial velocities (-50 to 50~km~s$^{-1}$). The gray shaded region shows the region of 1$\sigma$ uncertainty. Note that a broader range of radial velocities would extend the length of the solid black line, but the width of the 1$\sigma$ uncertainty would remain unchanged. The $UVW$s of moving groups come from \citet{torres08} and \citet{montes01}. The range of possible space motion for \sciencebin\does not overlap most of the known young moving groups with the exception of the Octans and Argus associations, but these associations are in a different part of the sky compared to\sciencebin\(Figure \ref{ykg}).}