2011

1 — 1101.1848

\caption[]{Highest ranked observation by \redc. Those observations with severe residual soft proton contamination have been excluded.}

\caption{\redc\versus\cratio. Red filled circles indicate those observations with time-variable SWCX signatures. Histograms of the \redc\and\cratio\values are shown in the side panels (using a bin size of 0.25).}

\caption{Lightcurve from comet C2001 Q4 (Neat) that resulted in the highest \redc\values (black - SWCX band, red - continuum band). The time axis is given in hours since the start of the\xmm\lightcurves. The solar wind proton flux as recorded by\ace\is given in blue. The\opluss\to\oplussix\ratio is plotted using the proton axis and is shown by the blue dashed line.}

\caption{Ratio of \mage/\oxys\to\oxye/\oxys\where available for the SWCX set. Where appropriate we mark the lower limit (red).}

\caption{Top: histogram of the modelled fluxes. Bottom: observed flux (0.25 to 2.5\,keV) versus the modelled flux for the SWCX set. A line (dashed, red) of gradient unity has been added to the graph to aid the eye.}

\caption{Example modelled (blue, in \fluns, left-hand y-axis) and \xmm\line-band (black,$\mathrm{ct}\,\mathrm{s}^{-1}$, right-hand y-axis) lightcurves. Observations with identifiers (top-left) 0150610101, (top-middle) 0054540501, (top-right) 0113050401 show the model lightcurve generally following the shape of the \xmm\lightcurve. Observations (bottom-left) 0141150101 and (bottom-middle) 0150320201 show the modelled lightcurve peak in a different period to the\xmm\lightcurve and (bottom-right) 0301410601 is an example from an observation without a SWCX enhancement. Five panels show the split between the SWCX-affected and SWCX-free periods (vertical dashed line).}

\caption{Example modelled lightcurve (blue, left-hand y-axis) with the \xmm\line-band (black, right-hand y-axis), for the case where the first eigenvalue percentage contribution was the lowest when comparing the modelled flux and\xmm\lightcurves. The contribution to the modelled lightcurve from the magnetosheath (green-dashed) and region past the bow shock (plum-dashed) are also shown. The SWCX-affected period was taken between the vertical dashed lines.}

\caption{Fractional difference between (top) the observed and modelled flux and (bottom) the observed and \modtwo\flux, versus the maximum solar wind flux. Also included in each panel is a histogram of the fractional differences. Cases where\xmm\is found within the helium focusing cone are marked in red.}

\caption{Histograms of mid-observation (left) GSE-X, (middle) GSE-Y and (right) GSE-Z \xmm\positions for good (blue) and bad (red) fractional differences between the observed and modelled fluxes. The histogram bins have been offset from one another in the plot, for ease of viewing.}

\caption[]{Table of the SWCX set observations, ranked by \redc\(the reduced-$\chi^{2}$ to the linear fit between the line-band and continuum lightcurves). Also listed for each case are the revolution number (Revn), observation (Obsn) and the MOS1 and MOS2 exposure identifiers (Expn M1 and Expn M2) and the ratio of the lightcurve variances (\cratio).}

2 — 1101.2160

\caption{XPS of Ti 2p core levels in Mn$_x$TiSe$_2$ ($x$=0.1 and 0.2) and their fits are presented. Spectral contributions of Ti2 and TI1 atoms are shown by a long dash line ( \broken) and short dash line (\dashed) respectively at both top and bottom frames of the graph. Corresponding Ti2/Ti1 ratio numbers are given.}

3 — 1101.2668

\caption{Zero-temperature, ohmic decay rate for the \textcolor{red}{$\bullet$ instantaneously coupled} and \textcolor{blue}{$\cdot$ gradually coupled} initial states of a two-level system with exponential cutoff frequency $\Lambda = 100 \, \Omega$. In this case the switch-on function is exponential, $\theta_\mathrm{s}(t) = 1-e^{-t/\tau_\mathrm{s}}$, and the switch-on times $\tau_\mathrm{s}$ are chosen to take the values $1/\Lambda$, $2/\Lambda$, $4/\Lambda$, $8/\Lambda$, $16/\Lambda$. }

\caption{Zero-temperature, ohmic decay rate for the \textcolor{red}{$\bullet$ unprepared} and \textcolor{blue}{$\cdot$ prepared} initial states of a two-level system with exponential cutoff frequency $\Lambda = 100 \, \Omega$. In this case preparation by freezing was used to create an initially excited state.}

4 — 1101.3318

\caption{The central line-of-sight velocity dispersion of our best-fitting, single-cluster models, i.e.\whose$r_{\rm c}$ is close to the one of Scl-dE GC$1$. All cases shown here have a SFE of $33$~\%. {\color{red}In the first two lines we give the results obtained in our non-DM simulations. The values of $r_{\rm c}$ in these models bracket that of the observed EGC - as discussed in the text, we expect that a good model of the observed data could be obtained for an intermediate value of the initial cluster Plummer radius.}}

5 — 1101.3641

\caption{(color online). Dynamical heterogeneities in the stress field. %Data were produced in the steady state and for When not otherwise specified, the system size is $N=2^{16}$. %Arrows in the graphs indicate a decrease of the strain rate ($\dot{\gamma}=10^{-2},10^{-2.5},10^{-3},10^{-3.5},10^{-4}$). {\it Top left}: Normalized two-time auto-correlation function of the stress as a function of strain ($\dot{\gamma}=10^{-2},10^{-2.5},10^{-3},10^{-3.5},10^{-4}$). The inset shows the same quantity as a function of time. {\it Top right}: Time evolution of the dynamical susceptibility (same values for $\dot\gamma$). {\it Bottom left}: Spatial shape of the normalized four point stress correlation function $G_{4}$ for a strain rate $\dot{\gamma}=10^{-4}$ at time $t^*$ where the dynamical susceptibility is maximal. {\it Bottom right}: Finite size scaling plot for the maxima of the dynamical susceptibilities. Shown are the maxima as a function of the inverse strain rate, both normalized by the system size ({\color{green} +} $N=2^{12}$, {\color{red} $\times$} $N=2^{14}$, $\ast$ $N=2^{16}$). The dotted line has slope one and the inset shows the raw data.}

\caption{(color online) {\it Left}: scaling of the diffusion coefficient $\tilde{D}$ with strain rate and system size ($\ast$ $N=2^{8}$, {\color{red} +} $N=2^{10}$, {\color{green} $\times$} $N=2^{12}$). The inset shows $\tilde{\mathcal{D}}$ versus strain $\gamma$ for $\dot{\gamma}=10^{-3}$. {\it Right}: Master curves for diffusion coefficient ($\times$) and dynamical susceptibility ({\color{red} +}), normalized by their plateau value as a function of strain rate in units of the corresponding critical strain rate, indicating the onset of finite size effects. The inset shows the unscaled master curves.}

6 — 1101.3987

\caption[]{The ionic $Pma2$ phase of ammonia is predicted to be stable above 90 GPa and consists of alternate layers of NH$_4^+$ and NH$_2^-$ ions. This view shows the three layers of the crystal structure. The top layer consists of NH$_2^-$ ions with orientation \begin{picture} (20,10) \put(1,0){\line(0,1){8}} \put(1,0){\line(1,0){8}} \end{picture} $\!\!\!\!\!\!\!$, the second layer consists of tetrahedrally bonded NH$_4^+$ ions and bottom layer consists of NH$_2^-$ ions with orientation \begin{picture} (20,10) \put(1,8){\line(1,0){8}} \put(1,8){\line(0,-1){8}} \end{picture} .}

7 — 1101.4131

\caption{Angular power spectrum~\eref{eq:spectrum} of the reconstructed Wigner function of \fref{fig:Wignerfunction}. Without damping ({\color{bulletcolor}\tiny $\boxempty$}) the power in modes $k\gtrsim70$ is too large and dominated by noise and aliasing effects; experimental uncertainties damp the angular power at large $k$ in a natural way ($\bullet$, see sections~\ref{sec:smooth} and~\ref{sec:smoothP}). Odd-$k$ modes contain less power than even-$k$ modes because of the approximate point symmetry of the Wigner function (see \fref{fig:Wignerfunction}).}

8 — 1101.5183

\caption{\label{fig:H_hop}\colorfig (a) The hopping parameters of the two-band hopping Hamiltonian $H_{\txt{hop}}$ in Ref.~\onlinecite{Ran:PRB:2009}; (b) The resulting Fermi surface for $t_1 = 1, t'_1 = 0.2, t_2 = 1.7,$ and $t'_2 = 0.3$ at half-filling (i.e., two electrons per site), where blue represents hole pockets while red represents electron pockets. The choice of parameters and filling fraction in (b) are assumed throughout this paper.}

\caption{\label{fig:phases}\colorfig The phase diagram of the half-filled orbital-symmetric two-band Hubbard Hamiltonian Eq.~\ref{eq:H_Hb}, where $J$ and $U$ are given in units of $t_1$ (for comparison, note that the full bandwidth of the noninteracting system is $W = 12.8~t_1$). Here P stands for paramagnetic metal, SDW stands for spin density wave metal, LM stands for antiferromagnetic insulator, and SL stands for spin liquid. The insulating region is indicated by the gray shade, and first-order phase boundaries are indicated by the yellow shade. For an explanation of the ``NNN decoupling'' and ``NN decoupling'' lines, see Fig.~\ref{fig:decoupling} and the main text.}

\caption{\label{fig:J_cuts}\colorfig The quasiparticle weight $Z$, the bond renormalization factor $Q^{f}_{ij}$ along the $x$, $y$, and diagonal ($x+y$) direction, and the orbital-diagonal staggered magnetization $m_0$ as function of $U$ for (a) $J = 0.4$, (b) $J = 0.8$, (c) $J = 1.2$ and (d) $J = 2.2$. The inset in (d) shows a blowup of the bond renormalization factors in the region where magnetization is enhanced.}

\caption{\label{fig:decoupling}\colorfig The real space picture of the renormalized hopping Hamiltonian on (a) the left of the ``NN decoupling'' line, (b) between the ``NN decoupling'' line and the ``NNN decoupling'' line, and (c) the right of the ``NNN decoupling'' line in Fig.~\ref{fig:phases}. Here the green dots denote the location of Fe atom, the violet (blue) lines denote non-zero renormalized nearest-neighbor hopping along the $x$ ($y$) direction, the brown lines denote non-zero renormalized next-nearest-neighbor hopping, and the gray dotted lines denote zero renormalized hopping.}

\caption{\label{fig:Sq}\colorfig Three patterns of renormalized bonds in a square lattice Hubbard model with non-mixing bands having only nearest-neighbor hopping. Here the green dots denote lattice sites, the blue line denote nonzero renormalized hopping, and the gray dotted lines denote zero renormalized hopping. It can be checked that within the slave-rotor mean field and on the insulating side (c) is energetically favorable to (b), which in turns is energetically favorable to (a).}

9 — 1101.5841

\caption{{a)~Generated photon flux in the Stokes ({\color{red} +}) and anti-Stokes ({\color{blue} o}) bands (1541.5-1558.5~nm and 1522-1528~nm, respectively), for $P=250~\mu$W. The solid curves are theoretical fits with a Bose-Einstein distribution (see Eq.\,\ref{eq:Pscat}). b)~Photon flux generated in the filtering line itself at Stokes ({\color{red} +}) and anti-Stokes ({\color{blue} o}) frequencies for input power $P=1.25$~mW. Solid curves: fit following the Raman noise in a silica fiber. c)~Photon pair flux generated in the silicon waveguide (+) and fit following eq.~\ref{eq:flux} (curve), input power: 1.25\,mW. Error bars are calculated from statistical error as well as error on out-coupling losses. In these figures, the flux for each data point is corrected by subtracting the loss spectrum and dark counts from the detectors. The experiments have been performed at different input powers to ensure similar statistical errors. The right axis (coloured in green) refers to the actual power while the left axis rescales the data to input power$P=250$\,$\mu$W to enable comparison with the spectrum reported in Fig.(a) (see discussion in main text).} }

\caption{Emitted flux response time. {Measured temporal profile of the input pump pulse ({\color{blue} *})} and corresponding total photon flux emitted at both Stokes and anti-Stokes frequencies (red - 3 top curves) for 300~$\mu$W~({\color{red} $\Delta$}), 1.25~mW~({\color{red} $\circ$}), and 2.5~mW~({\color{red} $\triangleleft$}) peak pump power. % The black dotted curve shows the temporal dynamic of the carrier density due to the laser pulse assuming a carrier lifetime of 1~ns. Rise/fall time ($10\%$-$90\%$) is the same (450~ps) for both the laser and the 3 scattered fluxes. The corresponding rise/fall time ($10\%$-$90\%$) of the carrier population is 2.3~ns assuming a 1~ns lifetime. The scattered flux is thus independent of the carrier population. }

10 — 1102.0669

\caption{Non-linear susceptibility. Imaginary part of the susceptibility as a function of the probe electric field $\mathcal{E}$ for the positive (a) and negative (b) scan directions for $\rho$ = 0.4 ({\color{violet}{$\opencircle\!$}}) 0.7 ({\color{red}{$\opensquare\!$}}) 1.0 ({\color{blu}{$\fullcircle\!$}}) 1.6 ($\fullsquare$) $\times10^{10}$~cm$^{-3}$. Data in (a) and (b) is fitted to third- and second-order non-linearities respectively, and the resulting density dependence plotted in (c) and (d). Both $\chi^{(2)}$ and $\chi^{(3)}$ display a quadratic density scaling, consistent with pair-wise interactions.\label{fig:chi}}

11 — 1102.4110

\caption{$X$ and $Y$ are genereted by adding together joint structure, individual structure, and noise. \textcolor{blue}{Blue} corresponds to negative values, \textcolor{red}{red} positive values.}

\caption{JIVE estimates for joint structure, individual structure, and noise. \textcolor{blue}{Blue} corresponds to negative values, \textcolor{red}{red} positive values.}

\caption{Sum of squared residuals in the simulated model (\textcolor{green}{green}), the fitted model with true ranks (\textcolor{blue}{blue}) and the fitted model under permutation testing (\textcolor{red}{red}) in 100 randomly generated simulations.}

\caption{Scatterplots of sample scores for the first two joint components, first two individual miRNA components, and first two individual gene expression components. Samples are colored by subtype: Mesenchymal (\textcolor{yellow}{yellow}), Proneural (\textcolor{blue}{blue}), Neural (\textcolor{green}{green}) and Classical (\textcolor{red}{red}). }

\caption{Plot of gene-miRNA correlations (A), and scores and loadings for the first two sparse joint components (B-E). In (A), gene-miRNA pairs are colored \textcolor{red}{red} if they have a significant positive correlation and \textcolor{blue}{blue} if they have a signficant negative correlation ($P<10^{-5}$). Panels (B) and (D) show sample scores for the first two joint components, colored by subtype. Panels (C) and (E) display gene-miRNA pairs where each have non-zero loadings. Pairs are colored \textcolor{red}{red} if both gene and miRNA loadings have the same sign, \textcolor{blue}{blue} otherwise. In panels (A), (C) and (E) genes and miRNAs are ordered separately by average linkage correlation clustering.}

12 — 1102.4624

\caption{\label{Vergleich}Ultraviolet fixed point in the $d$-dimensional Einstein-Hilbert theory. Comparison of universal eigenvalues $\theta'$, $\theta''$, $|\theta|$ and the invariant $\tau=\lambda_*(g_*)^{2/(d-2)}$ for different Wilsonian momentum cutoffs and various dimensions, normalised to the result for $R_{\rm opt}$ ($R_{\rm mexp}$ \textcolor{green}{$\bullet$}, $R_{\rm exp}$ \textcolor{red}{$\blacksquare$}, $R_{\rm mod}$ \textcolor{blue}{$\blacktriangle$}, $R_{\rm opt}$ $\blacktriangledown$); from \cite{Fischer:2006fz}.}

13 — 1102.4914

\caption{(a) The same data as in Fig.~1(b), but omitting that corresponding to the joint submission of Edinburgh and Heriot-Watt universities (which corresponds to the black disc) from the fitting procedure. (b) A comparison between statistics \& operational research (``\textcolor{red}{+}'' symbols and solid line (red online)) and applied mathematics (``\textcolor{blue}{$\times$}'' symbols and dashed line (blue online)). }

14 — 1102.5085

\caption{\coloronline{} The modular network representation~\protect{\cite{newman,newmanpark}}. (a) We obtain two networks by projecting onto elements or modules. (b) The failure of element 3 induces the failure of module B, uncoupling the remaining modules, even though the network itself remains connected. \label{fig:cartoon}}

\caption{\coloronline{} The size of the giant component $S$ for $r_m = \delta(m,\mu), s_n = \delta(n,\nu)$, with $\mu=3$ and $\nu=6$. Theory and simulations confirm that the network undergoes a transition from coupled to non-overlapping modules well before it loses global connectivity. Symbols represent the element ($\Square$) and module ($\Circle$) networks. \label{fig:results}}

\caption{\coloronline{} Robustness of scale-free networks. Here $r_m = \delta(m,3)$, $s_n \sim n^{-\lambda}$, $\fc = 1/2$, and $N \equiv \max \{n \mid s_n > 0\}$. Increasing $N$ and decreasing $\lambda$, measures known to improve the robustness of scale-free networks, actually magnifies the robustness gap. Surprisingly, this also increases the fragility of the module network, indicating that optimizing against structural failure may worsen the network's functional resilience. \label{fig:scalefree}}

15 — 1103.0166

\caption{(color online). $S_\perp(k)$ measured in units of $v_m^2$ in experiments {(where $v_m$ is the maximum velocity of the vibrating plate)}, for different values of $\phi$. Data represent averages over 3300 pairs of frames. The dotted lines on the right show the corresponding $T_g$. {As a reference for the eye, we also indicate the average (among the four experiments) of $T_b$, which is defined in the text.} Inset: $S_\rho(k)$ for $\phi=0.14$ ({\color{blue}$\blacktriangle$}) and $\phi=0.42$ ($\bullet$) with the analytical predictions for elastic disks~\cite{BC86}. In all figures the lowest wave-vector is $k\sigma \geq 2 k_{min}\sigma$ ($2 k_{min}\sigma = 4\pi\sigma/R=0.5$), because at smaller ones finite size-effect are dominant. The values of $S_\perp(k)$ reported are an average on moments with modulus $k \in [n 2\pi/R , (n+1) 2\pi/R]$ for integer $n \geq 2$.}

16 — 1103.1321

\caption{Exploded CAD view of one CHERCAM module, composed of $2 \times 2$ submodules of 16 PMTs inserted in a 15\,mm thick Ertalyte\textregistered \block, each with 64 bore holes of$25.7_{-0}^{+0.02}$\,mm diameter. The PMTs of one submodule are padded with rubber O-rings at their readout end and hold in place by the HV distribution board. Each submodule is clamped down using spacer screws which attach to 4 spacer rods. The spacer screws are in turn used to attach the front end readout board to the HV distribution board. This way the module becomes one stable unit of 64 PMTs ensuring their appropriate positioning in the array with a 27.5\,mm pitch at a tolerance of$\pm$0.02\,mm. The 3 rows of 4 long spacers, located along the submodule edges, are used to attach the module to the supporting grid.}

\caption{Result of the structural analysis when the instrument is subjected to a horizontal acceleration of 5\,g (yellow arrows). Zoom on the interface. The maximum Von Mises stress is reached in a fillet.}

17 — 1103.1402

\caption{Description of ppiTrim MITAB 2.6 columns \label{tbl:mitabcols}}{\footnotesize \rowcolors{2}{gray!25}{white} \begin{tabular}{rlp{5.0cm}>{\ttfamily\scriptsize}p{5.4cm}} \toprule Column & Short Name & Description & {\normalfont Example} \\ \midrule 1 & uidA & Smallest Gene ID of the interactor A$^{*\dagger}$ & entrezgene/locuslink:854647 \\ 2 & uidB & Smallest Gene ID of the interactor B$^*$ & entrezgene/locuslink:855136 \\ 3 & altA & All gene IDs of the interactor A$^*$ & entrezgene/locuslink:854647\\ 4 & altB & All gene IDs of the interactor B$^*$ & entrezgene/locuslink:855136\\ 5 & aliasA & All canonical gene symbols and integer CROGIDs of interactor A& entrezgene/locuslink:BNR1| icrogid:2105284\\ 6 & aliasB & All canonical gene symbols and integer CROGIDs of interactor B& entrezgene/locuslink:MYO5| icrogid:3144798\\ 7 & method & PSI-MI term for interaction detection method & MI:0018(two hybrid) \\ 8 & author & First author name(s) of the publication in which this interaction has been shown$^\ddagger$ & Tong AH [2002]|tong-2002a-3\\ 9 & pmids & Pubmed ID(s) of the publication in which this interaction has been shown & pubmed:11743162\\ 10 & taxA & NCBI Taxonomy identifier for interactor A & taxid:4932(Saccharomyces cerevisiae)\\ 11 & taxB & NCBI Taxonomy identifier for interactor B & taxid:4932(Saccharomyces cerevisiae)\\ 12 & interactionType & PSI-MI term for interaction type & MI:0407(direct interaction)\\ 13 & sourcedb & PSI-MI terms for source databases$^\ddagger$ & MI:0000(MPACT)|MI:0463(grid)| MI:0465(dip)|MI:0469(intact) \\ 14 & interactionIdentifier & A list of interaction identifiers$^\star$ & ppiTrim:tyuGkSOK231dh3YnSi6GbczJCFE=| MPACT:8233|dip:DIP-11198E|grid:147506| intact:EBI-601565|intact:EBI-601728| irigid:288990|edgetype:X \\ 15 & confidence & A list of ppiTrim confidence scores$^\bullet$ & maxsources:2|dmconsistency:full| conflicts:S3oaiXt5tA4vVrUsO1rc1TA9krk=\\ 16 & expansion & Either `none' for binary interactions or `bipartite' for subunits of complexes & none \\ 17 & biologicalRoleA & PSI-MI term(s) for the biological role of interactor A$^\ddagger$ & MI:0499(unspecified role) \\ 18 & biologicalRoleB & PSI-MI term(s) for the biological role of interactor B $^\ddagger$ & MI:0499(unspecified role) \\ 19 & experimentalRoleA & PSI-MI term(s) for the experimental role of interactor A$^\ddagger$ & MI:0496(bait)|MI:0498(prey)| MI:0499(unspecified role) \\ 20 & experimentalRoleB & PSI-MI term(s) for the experimental role of interactor B$^\ddagger$ & MI:0496(bait)|MI:0498(prey)| MI:0499(unspecified role) \\ 21 & interactorTypeA & PSI-MI term for the type of interactor A (either `protein' or `protein complex') & MI:0326(protein) \\ 22 & interactorTypeB & PSI-MI term for the type of interactor B (always `protein') & MI:0326(protein) \\ %23 & xrefsA & Not used by ppiTrim & - \\ %24 & xrefsB & Not used by ppiTrim & - \\ %25 & xrefsInteraction & Not used by ppiTrim & - \\ %26 & annotationsA & Not used by ppiTrim & - \\ %27 & annotationsB & Not used by ppiTrim & - \\ %28 & annotationsInteraction & Not used by ppiTrim & - \\ 29 & hostOrganismTaxid & NCBI Taxonomy identifier for the host organism & taxid:4932(Saccharomyces cerevisiae) \\ %30 & parametersInteraction & Not used by ppiTrim & - \\ 31 & creationDate & Date when ppiTrim was run & 2011/05/11 \\ 32 & updateDate & Date when ppiTrim was run & 2011/05/11 \\ %33 & checksumA & Not used by ppiTrim & - \\ %34 & checksumB & Not used by ppiTrim & - \\ 35 & checksumInteraction & ppiTrim ID for an interaction & ppiTrim:tyuGkSOK231dh3YnSi6GbczJCFE= \\ 36 & negative & Always `false' & false \\ \bottomrule \end{tabular} }

\caption{Randomly sampled deflated complexes from high throughput publications\label{tbl:smplcmpx1}}{\scriptsize \rowcolors{2}{gray!25}{white} \begin{tabular}{>{\ttfamily}lcrm{2.6cm}m{4.5cm}} \toprule{\normalfont\scriptsize ppiTrim Complex ID} & Sources & Pubmed ID & Members & Comments \\ \midrule 8AVRUHG76vkiFn2cZGICNZzr00Y= & grid & 14759368 & CFT2, YSH1, PTA1, MPE1 & Part of mRNA cleavage/polyadenylation complex (4/10 proteins). \\ 9yS57j/gbRbOlNmmimsVeonoraA= & grid & 14759368 & NUT1, MED7, MED4, SIN4, SRB4 & Part of mediator complex. \\ JU+EOkq6ipLh9DJKRtGRLUvT7vM= & grid,mint & 14759368 & UBP6, RPT3, RPN9, RPT1, RPN8, RPN2, RPN7, RPN1 & Part of proteasome. MINT does not contain complexes from the original paper. \\ HtTmhGiPyfIT2vFtRZ94uWw0rsY= & grid & 16429126 & IOC3, HTB1, HTA2, HHF2, ISW1, KAP114, ITC1, RPS4A, VPS1, NAP1, RPO31, ISW2, TBF1, BRO1, MOT1 & Part of Complex \# 99.\\ LnNzfyPGShcG7zkKynU6+fsK2eU= & grid & 16429126 & PSK1, NTH1, BMH2, RTG2, BMH1 & Part of complex \# 147 (two core proteins plus three attachments).\\ S2I6VRjFMWC6rkkM+oYXwKCg9YQ= & grid & 16429126 & RPL4B, MNN10, MNN11, HOC1, MNN9, ANP1 & Core complex (\# 111 -- mannan polymerase II) + one attachment protein (RPL4B).\\ 1fRmAapl2ruoQq202YUJg55maFo= & grid,mint & 16554755 & RSM24, RSM28, MRPS5, MRP13, MRPS35, RSM27, RSM7, RSM25, MRPS17, MRPS12, RSM19, MRP4 & Part of complex \# 1.\\ 5tBkYOmK/G1h3vaQmiOnUoBHHMQ= & grid,mint & 16554755 & CFT2, YSH1, MPE1, PAP1 & Part of complex \# 18.\\ 9f2DVj2rDGeCP53LHOnWRMwq14A= & grid,mint & 16554755 & KAP95, RTT103, VMA2, RAI1, RAT1, RPB2, SRP1 & True experimental association but not part of any derived complex. \\ AVawv51+6Fqe3DquygD/XfyrXxE= & grid,mint & 16554755 & RRP42, RRP45, RRP6, CSL4, MPP6, RRP4, LRP1, DDI1 & Part of complex \# 19.\\ NOLEwovavMsFrQEdkSUt/mldeMc= & grid,mint & 16554755 & CDC3, SHS1, CDC11, CDC12 & Part of complex \# 121.\\ WA51i87Lj1wGp/EeF1OV/YvbW1Y= & grid,mint & 16554755 & GTT2, TRX1, CRN1, SSA3, IPP1, CMD1, TRX2, TDH1, RPL40B, CDC21, OYE2 & True experimental association but not part of any derived complex.\\ YN/hQXQvzoB5HqrgPzVth28mGsY= & grid,mint & 16554755 & RRP43, RRP42, RRP45, RRP40, DIS3, RRP6, RRP4, LRP1 & Part of complex \# 19.\\ 1LRk+AgI8HpGOSAgkhDzNJWSvtI= & grid & 20489023 & RTG3, RTG2, TOR1, TOR2, CKA2, MYO2, MKS1, KOG1 & True experimental association. \\ xWzvxeJFGqjkCihjmQVf5gZhJjQ= & dip,grid,mint & 20489023 & PUF3, SAM1, GCD6, SPT16, MTC1, YGK3, LSM12 & True experimental association. \\ \bottomrule \end{tabular} }

\caption{Randomly sampled deflated complexes from low-throughput publications\label{tbl:smplcmpx2}}{\scriptsize \rowcolors{2}{gray!25}{white} \begin{tabular}{>{\ttfamily}lcrm{2.6cm}m{4.5cm}} \toprule{\normalfont\scriptsize ppiTrim Complex ID} & Sources & Pubmed ID & Members & Comments \\ \midrule 15VfQtoe5gxGNwPSY3AG0sq6A2U= & grid & 9891041 & CCR4, HPR1, PAF1, SRB5, GAL11 & NOT a true complex. This is because of bad annotation of PAF1--SRB5 interaction by the BioGRID. Completely opposite interpretation was given in the paper. \\ d79IdtwfTAENrH8CQ+c8CpS389Y= & grid & 10329679 & YPT1, VPS21, YPT7, GDI1 & True complex. This is the only experiment in the paper. \\ EtS4cgphEpTqJb/FS5qxyzf0ke8= & grid & 11733989 & CDC39, CCR4, CDC36, CAF130, CAF40, CAF120, POP2, NOT5, MOT2 & True complex. CAF120 is an unusual member that could almost be left out. \\ 2kOyGdwzWywSpN5mhK26gCcC6LQ= & grid & 14769921 & GBP2, IMD3, TEF1, KEM1, CTK2, CTK1, CTK3 & True complex, except that TEF1 should be TEF2. This is an error in the iRefIndex source file; the BioGRID website has the correct assignment. \\ Kd07BBUF07Sqy9NP3D0lixsS/TY= & grid & 15303280 & BUD31, RPL2B, PRP19, CDC13, ATP1, RPS4A, SNU114, MDH1, MAM33, MRPL3, MRPL17, PRP8, PRP22, PAB1, BRR2 & True association \\ ZAGz/IZqkEr3/NTDLzPEDAD9cKo= & grid & 16179952 & CDC40, UFD1, SSM4, UBX2 & NOT a true complex, probably due to a typo in annotation. CDC40 cannot be found anywhere in the paper and should most likely be CDC48. \\ RDu0dsPAN0QEadfSU5sv05Ifihw= & grid & 16286007 & SIN3, RCO1, RPD3, UME1, EAF3 & True complex. \\ Vqbn3dDwTPgyE9DzbatFNqzdFe0= & grid & 16615894 & VPS36, VPS25, VPS28, SNF8 & Vps28 binds the other three, which form a complex. \\ lmdypAN9kaHBdasLWS19x8K7KkE= & grid & 20159987 & UBI4, UFD2, PEX29, SSM4 & Biological association but indicated as `NOT a stable complex' in the paper. \\ aakRh6qVahGxGvqHe399+faxPvA= & grid & 20655618 & PEX13, PEX10, PEX8, PEX12 & Association is correct, although mutant strain was used to obtain this particular complex. \\ \bottomrule \end{tabular} }

\caption{Summary of resolvable conflicts \label{tbl:resolvable_sce}}{\scriptsize \rowcolors{2}{gray!25}{white} \begin{tabular}{>{\raggedright}p{12cm}r} \toprule Consolidated terms & Count \tabularnewline \midrule MI:0018 (two hybrid), MI:0045 (experimental interaction detection), MI:0398 (two hybrid pooling approach), MI:0399 (two hybrid fragment pooling approach) & 3959 \tabularnewline MI:0090 (protein complementation assay), MI:0111 (dihydrofolate reductase reconstruction) & 2612 \tabularnewline MI:0090 (protein complementation assay), MI:0112 (ubiquitin reconstruction) & 2077 \tabularnewline MI:0004 (affinity chromatography technology), MI:0676 (tandem affinity purification) & 1840 \tabularnewline MI:0004 (affinity chromatography technology), MI:0007 (anti tag coimmunoprecipitation) & 1408 \tabularnewline MI:0018 (two hybrid), MI:0045 (experimental interaction detection), MI:0397 (two hybrid array) & 1231 \tabularnewline MI:0018 (two hybrid), MI:0045 (experimental interaction detection) & 954 \tabularnewline MI:0018 (two hybrid), MI:0397 (two hybrid array) & 914 \tabularnewline MI:0045 (experimental interaction detection), MI:0686 (unspecified method) & 628 \tabularnewline MI:0004 (affinity chromatography technology), MI:0019 (coimmunoprecipitation) & 598 \tabularnewline MI:0018 (two hybrid), MI:0398 (two hybrid pooling approach) & 506 \tabularnewline MI:0004 (affinity chromatography technology), MI:0007 (anti tag coimmunoprecipitation), MI:0676 (tandem affinity purification) & 444 \tabularnewline MI:0018 (two hybrid), MI:0045 (experimental interaction detection), MI:0686 (unspecified method) & 320 \tabularnewline MI:0004 (affinity chromatography technology), MI:0096 (pull down) & 217 \tabularnewline MI:0415 (enzymatic study), MI:0424 (protein kinase assay) & 192 \tabularnewline MI:0045 (experimental interaction detection), MI:0081 (peptide array) & 150 \tabularnewline MI:0045 (experimental interaction detection), MI:0676 (tandem affinity purification) & 120 \tabularnewline \midrule MI:0492 (in vitro), MI:0493 (in vivo) & 5739 \tabularnewline MI:0018 (two hybrid), MI:0398 (two hybrid pooling approach) & 5394 \tabularnewline MI:0018 (two hybrid), MI:0492 (in vitro), MI:0493 (in vivo) & 2796 \tabularnewline MI:0096 (pull down), MI:0492 (in vitro), MI:0493 (in vivo) & 2760 \tabularnewline MI:0096 (pull down), MI:0492 (in vitro) & 2134 \tabularnewline MI:0018 (two hybrid), MI:0492 (in vitro) & 1658 \tabularnewline MI:0018 (two hybrid), MI:0493 (in vivo) & 1193 \tabularnewline MI:0018 (two hybrid), MI:0397 (two hybrid array) & 1045 \tabularnewline MI:0096 (pull down), MI:0493 (in vivo) & 513 \tabularnewline MI:0004 (affinity chromatography technology), MI:0006 (anti bait coimmunoprecipitation) & 384 \tabularnewline MI:0004 (affinity chromatography technology), MI:0019 (coimmunoprecipitation) & 309 \tabularnewline MI:0004 (affinity chromatography technology), MI:0007 (anti tag coimmunoprecipitation) & 195 \tabularnewline MI:0114 (x-ray crystallography), MI:0492 (in vitro) & 166 \tabularnewline MI:0004 (affinity chromatography technology), MI:0096 (pull down) & 161 \tabularnewline MI:0047 (far western blotting), MI:0492 (in vitro), MI:0493 (in vivo) & 106 \tabularnewline \midrule MI:0018 (two hybrid), MI:0398 (two hybrid pooling approach) & 17738 \tabularnewline MI:0018 (two hybrid), MI:0399 (two hybrid fragment pooling approach) & 1426 \tabularnewline \bottomrule \end{tabular} }

18 — 1103.3521

\caption{\textcolor{green}{\label{fig:damping-times}}The relevant r-mode time scales for $1.4\, M_{\odot}$ stars rotating at their Kepler frequency. \emph{Left panel: }Neutron star. \emph{Right panel: }Strange star. The dotted horizontal line presents the time scale $\tau_{G}$ associated to the growth of the mode due to gravitational wave emission. The dashed rising curve shows the damping time $\tau_{S}$ due to shear viscosity. The damping time $\tau_{B}$ due to bulk viscosity is given for different dimensionless r-mode amplitudes $\alpha=0$, $0.01$, $0.1$, $1$ and $10$ by the solid curves. The thin dotted curves correspond to the analytic linear approximation eq. (\ref{eq:linear-bulk-vicosity-time}) and are below the shown plot range for the largest amplitude. The thin dot-dashed curves on the left panel show the change when only a smaller core (ranging to a density of $n_{0}/2$ instead of $n_{0}/4$) is taken into account. The thin dashed curves on the right panel represent the approximate analytic expression eq. (\ref{eq:homogenous-quark-damping-time}) given in the appendix which is not valid above the maximum of the bulk viscosity and therefore not shown for the large amplitude results.}

19 — 1103.3658

\caption{(a) Magnetization of the filament $m=(\rhobp-\rhobm)/(\rhobp+\rhobm)$ and (b) current of positive particles along both filaments in the canonical setup with a static filament, for a system size $L=200$ and different densities $\rho$. Full lines (\full) are for densities $\rho<1$ and dashed lines (\dashed) are for densities $\rho\geq1$. All quantities are plotted against the interaction strength $q$ tuning the modified attachment/detachment rates.}

20 — 1103.5917

\caption{{\bf Genome wide and regional sex-differences in recombination rates.} {\bf A)} The difference between female and male recombination rates ( $\circ$ from linkage maps, excluding known sex-chromosomes, $\times$ from chiasmata counts) divided through by the sex-averaged rates. Points above the dashed line indicate higher rates of recombination in females than in males. $^*$ indicates $p < 0.05$ using a two-tailed sign test, without correcting for multiple tests or phylogeny, and ignoring ties. {\bf B)}. Sex-standardized recombination rates across the human genome. The sex-standardized rate equals the local recombination rate in a given sex (\textcolor{blue}{male} and \textcolor{red}{feamle}), divided by the average recombination rate in that sex. The x-axis indicates the position of the focal genomic region (.2\% of a chromosome arm), divided by the length of the chromosome arm. Data are presented from all metacentric human autosomes. Lines represent a lowess smoothing of these points. }

\caption{{\bf The coevolution of an MI driver and an unlinked recombination enhancer.} The frequencies of MI drive alleles \textcolor{red}{($f_D$, red)}, and unlinked recombination modifiers {\color{SkyBlue}($f_M$, blue)} across generations. The correlation between alleles, $\frac{LD}{\sqrt{f_Df_df_Mf_m}}$, denoted by the \textcolor{red}{red}{\color{SkyBlue} blue} line, and its value is given on the right axis. Drive is complete and recessive lethal ($w_{Dd}=1$, $w_{dd}=0$). The initial recombination rate is 1/4, and each copy of M increases the probability of recombining by 0.05. Initial frequencies of drive and recombination modifier alleles equal $f_{D_0}=0.10$ and $f_{M_0}=0.01$, respectively. {\bf \ref{Unlinked}A)} Drive in both sexes ($\alpha_{\text{MI}}=1$, $r=1/4$, $\delta r = 0.05$). {\bf \ref{Unlinked}B)} Female-limited drive ($\alpha_{\text{MI}\Venus}=1$, $\delta r_\Venus = 0.05$). }

\caption{ {\bf The evolution of drivers and recombination modifiers in tight linkage} The frequencies of MI drivers \textcolor{red}{($f_D$, red)}, and linked recombination modifiers {\color{SkyBlue}($f_M$, blue)} across generations. The correlation between alleles is denoted by the \textcolor{red}{red}{\color{SkyBlue}blue} line, and its value is given on the right axis. Initial frequencies of driver and recombination modifier alleles are $f_{D_0}=0.10$ and $f_{M_0}=0.01$, respectively. Drive is complete and recessive lethal. Initial recombination rate equals 1/4. {\bf \ref{Linked}A}) Drivers and recombination enhancement in both sexes ($ \delta r = 0.05$), M arises on a d chromosome. {\bf \ref{Linked}B}) Female-limited driver and recombination enhancement. ($ \delta r_\Venus = 0.05$), M arises on a d chromosome. {\bf \ref{Linked}C}) Drivers and recombination suppression in both sexes. ($ \delta r = - 0.05$), M arises on a D chromosome. {\bf \ref{Linked}D}) Female-limited driver and recombination suppression. ($\delta r_\Venus = -0.05$), M arises on a D chromosome. }

21 — 1104.0146

\caption{(a) The~$x$--$y$ plane divided into five regions. The thick black line corresponds to the wall. (b) An example of a pie chart. The different \protect\ifcolor{colors}{shadings} as a function of initial orientation~$\theta_0$ indicate which region of (a) the treadmilling organism ends up in after a sufficiently large time.}

\caption{Pie charts at $t=1500$ for several initial conditions. Each pie chart is centered on the initial condition it corresponds to. The thick black line represents the wall. \protect\ifcolor{Note}{To help distinguish the shadings, note} that region 3 only appears as very thin slivers in the pie charts along the $x$ axis.}

22 — 1104.2704

\caption{(Colour online) Phasespace of the pendulum (black \dashed) and the kicked rotor (grey \full) for $\tilde{k}=|\epsilon| k=0.08\pi$.}

\caption{(Colour online) Fidelity using the pendulum (green \dashed), the perturbative result with the third (black \full) and fourth (blue \dashddot) order in $|\epsilon|$ in the energy, and the original QKR (red \chain). $\beta=0.3$, $\epsilon=0.05$, $k_1=0.6\pi$ and $k_2=0.8\pi$. The data are averaged over $100$ kicks in order to cancel fast oscillations.}

\caption{(Colour online) Fidelity for the QKR (red~\full) and the pendulum (green~\dashed) for $\epsilon=0.075$, $k_1=0.6\pi$ and $k_2=0.8\pi$. (a) $\beta=0.3216$, (b) $\beta=0.2412$. The data are averaged over $100$ kicks in order to get rid of the fast oscillations.}

\caption{(Colour online) Fidelity of a few ensembles of $\beta-$rotors with $\epsilon=0.05$, $\Delta\beta=0.06$, $k_1=0.6\pi$, $k_2=0.8\pi$, $\beta_1=0.08$ (black \full), $\beta_1=0.12$ (red \dotted), $\beta_1=0.16$ (green \dashed), $\beta_1=0.2$ (blue \chain) and $\beta_1=0.24$ (magenta \dashddot). In (a) the original data and in (b) the rescaled data are shown. The scaling factors are according to figure~\ref{fig:Scaling_factrors}.}

23 — 1104.3616

\caption{Performance comparison of strategic trading (real data) and random trading using the average values of return $R$ versus trading frequency $J$. We exclude the sell transactions without any preceding matching buys. The simulations for the random strategies are repeated for 2000 times. We show the results for individuals in A-share (a), institutions in A-share (b), individuals in B-share (c) and institutions in B-share (d), respectively. In each plot, the colorful symbols ({\color{black}{$\circ$}}, {\color{red}{$\vartriangle$}}, {\color{green}{$\triangledown$}}) correspond to strategic trading, the continuous lines correspond to random trading, and the dashed line indicates the base line of zero return ($R=0$).}

\caption{Average returns $R$ versus average holding time for individuals in A-share (a), institutions in A-share (b), individuals in B-share (c) and institutions in B-share (d), respectively. The symbols present the average values over all investors ($\circ$), as well as investors who earn positive return ({\color{red}{$\vartriangle$}}) and negative return ({\color{green}{$\triangledown$}}). The dashed line delineates the benchmark in absolute terms of zero return. The solid lines with error bars are the average simulation results of random trading. The insets in panels (c) and (d) are magnifications of the curves for small values of $\Delta t$.}

24 — 1104.3844

\caption{ Holevo phase variance $V_\text{H}$ of PSO optimized GLS policies. Purple crosses (\textcolor{maroon}{\protect\scalebox{1.2}{$\times$}}): $N$-qubit policy used with input state $\ket{\Psi_N}$. Brown pluses (\textcolor{brown}{\protect\scalebox{0.8}{$\bm{+}$}}): $(N-1)$-qubit policy used with input state $\ket{\Psi_N}$ (the last measurement result is ignored by the policy). }

\caption{ Holevo phase variance $V_\text{H}$ of policies optimized for simulated Gaussian quantum noise (\textcolor{green}{$\bullet$}) and skew-normal quantum noise with skewness $\gamma = 0.667$ (\textcolor{blue}{$\bullet$}). In both cases, we used the standard deviations $\sigma_\theta = 0.02\pi$ and $\sigma_{\bm{n}_x} = \sigma_{\bm{n}_y} = \sigma_{\bm{n}_z} = 0.2\sigma_\theta$. }

\caption{ Holevo phase variance~$V_\text{H}$ of policies from \cite{QLearning:hentschel:PRL:2010}, that are optimized for a perfect interferometer (\textcolor{blue}{$\Diamond$}), compared to the policies optimized by our new algorithm for the specific imperfections (\textcolor{darkred}{$\bullet$}). The performance of the policies are evaluated for Gaussian quantum noise with standard deviations $\sigma_\theta$ and $\sigma_{\bm{n}} = (\varepsilon, \sqrt{1- \varepsilon^2},\varepsilon)$, $\varepsilon = 0.2\sigma_\theta$. (a) For low noise ($\eta = 5\%$ and $\sigma_\theta = 0.02\pi$), there is no noticeable performance enhancement. (b) For larger noise ($\eta = 5\%$ and $\sigma_\theta = 0.1\pi$), the policies optimized for perfect interferometry have a performance scaling $V_\text{H} \propto N^{-\alpha}$ with $\alpha_\tn{1} = {1.162 \pm 0.003}$. In contrast, the policies optimized for the aforementioned noise and loss achieve a scaling of $\alpha_\tn{2} = {1.236 \pm 0.003}$. }

25 — 1104.3868

\caption{Mass accretion rates onto the BHs, for the runs with $e_{0}= 0.2,0.4,0.6,0.8$ (bottom to top). Dashed (\textcolor{red}{red}) line refers to the primary BH, solid (\textcolor{blue}{blue}) line to the lighter secondary hole.}

\caption{Power spectrum of the accretion rate (in arbitrary units$^2$) onto the primary (\textcolor{blue}{blue}) and secondary (\textcolor{red}{red}) BHs. Frequencies are in units of the initial binary orbital frequency $f_0$. The inlays show zoom-ins of the power spectra, in the frequency range $0.1$--$1 f_0$ computed summing the accretion rate from the two BHs (\textcolor{RubineRed}{pink}). The expected intervals for the disc and the beat frequencies are marked by the thick vertical black lines, as labelled in the Figure. }

26 — 1104.4337

\caption{ Comparison of available alignment strategies for nucleic acid CE profiles. (A) Comparison of electrophoretic profiles of the 88-profile MedLoop DMS mutate-and-map data set~\citep{kladwangcordero2011} (replicate 2) before alignment and after alignment by HiTRACE, msalign~\citep{kazmi2006alignment}, SpecAlign~\citep{wong2005specalign}, COW~\citep{tomasi2004correlation} and CAFA~\citep{mitra2008high}. (B) Alignment results for the 480-profile P4-P6 DMS data set. {\color{red}(C) Quantitative comparison of alignment results for all 13 data sets based on mean squared error (MSE;~\citealp{kay1993fundamentals}) of aligned peak positions with respect to the reference peaks. Red line is median value; box boundaries represent 75th and 25th percentiles; error bars represent the most extreme values whose distance from the box is less than 1.5 times the box length; + symbols are outliers beyond this range. (D) Quantitative comparison in terms of the Kullback-Leibler divergence~\citep{cover2006elements} between reference and non-reference profiles.} }

\caption{ \color{red}Quantification accuracy and precision for HiTRACE and CAFA. Correlation of HiTRACE and CAFA results for two MedLoop DMS data sets [(A)~\&~(B)] confirms the absence of any systematic deviation between the two approaches. Precision of HiTRACE [(C)~\&~(D)] is similar or better than CAFA [(E)~\&~(F)], based on independent analyses of the same data set. }

\caption{\textcolor{red}{Correlation of results between experimental replicates for HiTRACE (A) and CAFA (B) for the MedLoop DMS mutate-and-map experiments~\citep{kladwangcordero2011}, and (C) for HiTRACE on five replicate data sets without (black bars) and with (white bars) optimization of Gaussian positions during band deconvolution.}}

\caption{The average running time of HiTRACE on different data sets. The time was measured on a personal computer system equipped with a \textcolor{red}{2.66GHz Core i5 processor (4 cores; multithreading enabled)} with 4GB RAM. }

\caption{Overview of the proposed HiTRACE methodology. { \color{red} (A) Raw electropherograms for an example data set. (left) Dimethyl sulfate (DMS) modification of the MedLoop RNA~\citep{kladwangcordero2011}, read out by reverse transcription with rhodamine-green-labeled primers followed by DNA separation by capillary electrophoresis; data shown are DMS profiles for 80 (of 120) single nucleotide mutants, two replicate controls without chemical modification, and sequencing ladders for C, U, and G. (right) Electropherograms of the Texas-red-labeled DNA ladder that was co-loaded with each sample to produce fiducial markers for alignment. (B) Profiles after automated preprocessing (baseline subtraction) and correlation-optimized linear alignment. (C) Profiles after automated alignment refinement by dynamic-programming-based nonlinear adjustments. (D) Interactive sequence annotation guided by features (red circles) at mutation positions and bands in the sequencing ladder. Blue, cyan, orange, and red lines correspond to modifications at A, C, G, and U, respectively. (E) Quantitated band areas (the final output of HiTRACE). Shorter DNA fragments (higher mobility) are at the top of each panel.} }

27 — 1104.5021

\caption[fig:Diffusion]{(color online). Evolution under discrete quantum walks for $N=4$. The symbols represents different types of states: `\textcolor{PineGreen}{\textbf{S}}' for the initial state, `\textcolor{orange}{\textbf{X}}' for states explored by a quantum walk, `\textcolor{blue}{\textbf{\#}}' for unexplored target states, `\textcolor{red}{\textbf{@}}' for explored target states, `\textcolor{black}{\textbf{.}}' for remaining unexplored functional basis states. At least $N=4$ steps are necessary to reach the target states, which is an illustration of Eq.~(\ref{eq:QW}). (a) For UDD, a discrete quantum walk occurs in a one-dimensional functional basis, with initial state $\left\vert 1\right\rangle $, target state $\left\vert N+1\right\rangle $, and quantum walk matrix $D_{1}$. (b) For QDD, a discrete quantum walk occurs in a two-dimensional functional basis, with initial state $\left\vert 11\right\rangle $, target states $\left\{ \left\vert \kappa ^{\#}\right\rangle \right\} $, and a quantum walk matrix $D_{\left( 01\right) }$.}

28 — 1104.5041

\caption{\label{fig:schematic}Schematic diagram of the magnetically coupled piston pump. One of the reed valve blocks, made from 4-5/8\texttt{"} ConFlat\textsuperscript\textregistered\flanges, is shown separately at the top right.}

29 — 1104.5378

\caption{\label{fig:phase_diagrams} \emph{Phase diagrams} for (a) a single random walker on a periodic lattice with varying length $L$ and (b) varying number of random walkers and fixed $L=160$. Dashed lines represent the phase boundary calculated from the FB picture, eq.~\eqref{manyb:cond}. Solid line is a guide to the eye indicating the actual phase boundary. Employing a Gillespie algorithm, the parameter regime has been scanned with a resolution of $\delta r_\pm = 0.05$. Red crosses ({\color{red}+}) indicate a phase with positive velocity, blue crosses ({\color{blue}$\times$}) a phase with negative velocity, and green stars ({\color{green}$\star$}) a phase with current reversal.}

30 — 1105.0264

\caption{Convergence with respect to the plane-wave cutoff energy $\varepsilon_\textrm{cut}$ for the calculated loss function $-\Im\{\varepsilon^{-1}(\textbf{q},\omega)\}$ versus energy $\omega$ in~eV and momentum transfer $\|\textbf{q}\|$ in~\AA$^{-1}$ along the armchair direction of graphene, including $n_{\textrm{unocc}}$ = 8 bands/atom, and a radial cutoff $R =$ 4 \AA. Results for $\varepsilon_\textrm{cut}$ = 15 ({\color{blue}{$\cdots \cdots$}}), 30 ({\color{red}{-- -- --}}), 60~eV (------), and 90~eV ({\textbf{------}}), corresponding to 11, 27, 89, and 149 $\textbf{G}$ vectors respectively, are shown.}

31 — 1105.0923

\caption{Plot of the X-ray and radio luminosities of all YSOs that have been simultaneously detected in both the X-ray and radio ranges by \textsl{Chandra} and the VLA, excluding sources with a radio detection significance of $<5\sigma$ and less than 100 counts in X-rays. The symbols `+', `$\times$', and `\textasteriskcentered' indicate YSO classes I, II, and III, respectively. Apart from sources reported in this paper, and sources in $\rho$ Oph from \citet{gag04}, scaled to a distance of 130~pc, IRS sources are in CrA \citep{for07}, and L1 stands for LkH$\alpha$\,101\citep{ost09}. The central line marked ``GB'' indicates the G\"udel-Benz relation with its upper and lower bounds, as reported by\citet{gue02}, shown as dashed lines. The vertical lines indicate the 5$\sigma$ sensitivities of the different observations, as converted into radio luminosities at the respective distances. The limits are strictly for the simultaneous observations only (some regions have deeper radio and/or X-ray data from separate observations). The X-ray sensitivity limits are lower than the X-ray luminosity range shown (see Table~\ref{tbl-cxo-vla}).\label{n1333_ic348_xr_plot2}}

32 — 1105.0992

\caption{(Color online) The conductance $G$ vs bar width $L_y$ at disorder strength $W=115m$eV. Different lengths are chosen with $L_{x}$=1000nm (black solid line), 2500nm (red dash line), 4000nm (green dot line), and 5000nm (blue short dash line), while the symbols {\tiny\textcolor{black}{$\blacksquare$}}, {\tiny\textcolor{red}\textbullet}, {\tiny\textcolor{green}{$\blacklozenge$}} and {\tiny\textcolor{blue}{$\blacktriangle$}} are corresponding to the realistic results of exact diagonalization, respectively. The inset shows the linear fitting of ln(2-G) by picking a typical length $L_x=5000nm$.}

33 — 1105.1051

\caption{\coloronline Joint probability distribution of two particles after $t=50$ steps of a Hadamard walk (antisymmetric subspace). Left panel: without interaction ($\gamma=1$); Right panel: interaction with collision phase $\gamma=-1$. The embedded (green) squares are for comparison with the theoretically calculated peak velocity of the free Hadamard walk (left panel) and the peak velocity of the effective walk of molecules after Eq.~\eqref{velo}. }

\caption{\coloronline Spectrum of the interacting walk operator for two particles on a ring (28 sites). Left Panel: The non-interacting case: The relative momentum is then a good quantum number, and is represented as a third direction, orthogonal to the paper plane. For large rings the points fill the indicated surfaces. Right Panel: The interacting case ($\gamma=-1$). The bands are very nearly equal to the projection of the left-panel figure onto the $(p,\omega)$-plane. The additional feature introduced by interaction is the chain of simple eigenvalues in the band gap, i.e., the molecule states.}

34 — 1105.2174

\caption{Transverse MR at: 30 ($\blacksquare$), 40 (\textcolor[rgb]{1,0,0}{$\bullet$}), 50 (\textcolor[rgb]{0,0,1}{$\blacktriangle$}), 60 (\textcolor[rgb]{0,0.5,0.5}{$\blacktriangledown$}), 80 (\textcolor[rgb]{1,0,1}{$\blacktriangleleft$}), 120 (\textcolor[rgb]{0.5,0.5,0}{$\blacktriangleright$}) and 200 K (\textcolor[rgb]{0,0,0.5}{$\blacklozenge$}) in transverse magnetic fields. a) MR increased monotonically with decreasing temperature. In the strong magnetic field condition MR was non-saturating and approximately linear. b) MR at all temperatures reduced down to one unique curve provided $\mu B = \omega_{c} \tau$ conforming to Kohler's rule. The data is magnified in the inset to demonstrate the weak to strong field crossover at $\omega_{c}\tau \sim 1$ and universal scaling of the weak field MR. A quadratic law is plotted at weak fields, $\omega_{c}\tau \ll 1$ (dashed line)}

\caption{Geometrical enhancement MR, measured in perpendicular (a) and transverse (b) field orientations at 35 K shown mirrored to one another using the right hand axis to encourage comparison. Wafers A-H are shown with respective values of $\alpha = l / w$: $10 \pm 1 $ (A,$\blacksquare$), $4.5 \pm 0.6$ (B,\textcolor[rgb]{1,0,0}{$\bullet$}), $3.7 \pm 0.5$ (C,\textcolor[rgb]{0,0,1}{$\blacktriangle$}), $2.6 \pm 0.3$ (D,\textcolor[rgb]{0,0.5,0.5}{$\blacktriangledown$}), $2.3 \pm 0.2$ (E,\textcolor[rgb]{1,0,1}{$\blacktriangleleft$}), $1.9 \pm 0.3$ (F,\textcolor[rgb]{0.5,0.5,0}{$\blacktriangleright$}), $1.6 \pm 0.2$ (G,\textcolor[rgb]{0,0,0.5}{$\blacklozenge$}), $1.1 \pm 0.2$ (H,\textcolor[rgb]{0.5,0,0}{$\bigstar$}).}

35 — 1105.2593

\caption{(Color online) Band structure (lines) and energies of the surface and subsurface states shown in Fig.~\ref{Fig2} (symbols) for the B terminated MgB$_2$(0001) surface in eV relative to the Fermi level $\varepsilon_F$. Results from plane wave PWscf calculations before ({\color{blue}{\bf{---}}}; {\color{blue}{$\blacksquare$}}) and after ({\color{red}{\bf{-- --}}}; {\color{red}{$\medbullet$}}) structural optimization are compared with real-space GPAW calculations for the relaxed structure ($\cdots$; $\vartriangle$).}

\caption{(Color online) GPAW overlap-based band structure of relaxed B terminated MgB$_2$(0001) surface, with \(\psi_{n,\mathbf{k}}\rightarrow\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\) for \(n'\) which maximizes \(\left|\langle\psi_{n,\mathbf{k}}|\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\rangle_{\mathcal{V}}\right|\). Surface states with more than two thirds of their weight in the first vacuum and first two bulk layers, $s_{n,\mathbf{k}} \gtrsim 0.66$, are shown separately ({\color{red}{\bf{---}}}).} \label{Fig_GPAW_B} \end{figure} The DFT estimates of the surface formation energy $\Delta E_f$ shown in Table~\ref{GPAW_DEf} are quite small ($|\Delta E_f| \lesssim 0.05$~eV/\AA$^2$), which suggests that each of the MgB$_2$(0001) surface terminations considered should be stable experimentally. This is clear when one compares with the much higher surface formation energies obtained for other materials, such as metal oxides, which are known to be stable experimentally \cite{Trends}. In particular, we find it should be thermodynamically preferred to replace a Mg surface termination by a Li termination, so that surface doping by Li is quite feasible. It should be noted that our choice of bulk metallic B, Mg, and Li for the chemical potential's shown in Table~\ref{GPAW_DEf} has a significant influence on the resulting formation energies obtained. However, the main conclusion that each surface termination should be obtainable under particular experimental conditions should still hold. The projection of the bulk band structure in a direction perpendicular to the crystal surface ($\Gamma\rightarrow\mathrm{A}$) is a fingerprint of the particular crystal structure. The same is true for a MgB$_2$(0001) surface, for which the projected band structure is actually quite simple. Due to the weak B--B layer overlap, there are three rather narrow B $\sigma$ bands ($\sigma_1$, $\sigma_2$, and $\sigma_3$), while the overlap between $p_z$ orbitals in the B layers and $s$ orbitals in Mg layers yields a wider $\pi$ band. The $\pi$ band is then separated from $\sigma$ bands by two wide gaps, which are noticeable in Figs.~\ref{Fig1}, \ref{Fig3} and \ref{Fig5}. Since the surface band structure of B and Mg terminated MgB$_2$(0001) surfaces was previously studied in Refs.~\onlinecite{Slava,Profeta}, we will focus here on the detailed classification of the surface and subsurface states and surface state resonances in Mg, B and Li terminated MgB$_2$(0001). This is accomplished by combining both pseudopotential-based plane wave methods (SPPW, PWScf) and a real space projector augmented wave function method (GPAW) to test the robustness of the surface states, and geometry optimization to show how surface relaxation can modify them. \subsection{B-terminated MgB$_{\bf 2}$(0001) surface} \begin{table} \caption{Change in B terminated Mg--B inter-layer separation $\Delta_{i,j}$ between the $i^{\mathrm{th}}$ and $j^{\mathrm{th}}$ layers from the surface, relative to the bulk MgB$_2$ experimental geometry.}\label{Table1} \begin{ruledtabular} \begin{tabular}{lrrrr} &\multicolumn{2}{c}{PWscf LDA\footnote{This Work}}&\multicolumn{2}{c}{VASP PBE\footnote{Ref.~\onlinecite{Li}}} \\\hline $\Delta_{0,1}$[B--Mg]& $-4.4\%$&-7.8 pm& $-2.1\%$& -3.7 pm\\ $\Delta_{1,2}$[Mg--B]& $ 0.4\%$& 0.75 pm& $ 2.0\%$& 3.5 pm\\ $\Delta_{2,3}$[B--Mg]& $-1.1\%$&-2.0 pm& $ 0.9\%$& 1.6 pm\\ $\Delta_{3,4}$[Mg--B]& $-1.7\%$&-3.1 pm& $-1.8\%$& -3.2 pm\\ $\Delta_{4,5}$[B--Mg]& $-1.3\%$&-2.6 pm& $0\%$& 0.0 pm\\ $\Delta_{5,6}$[Mg--B]& $-1.3\%$&-2.3 pm&---&---\\ $\Delta_{6,7}$[B--Mg]& $-1.5\%$&-2.7 pm&---&---\\ $\Delta_{7,8}$[Mg--B]& $-1.3\%$&-2.3 pm&---&---\\ $\Delta_{8,9}$[B--Mg]& $-0.75\%$&-1.4 pm&---&---\\ \end{tabular} \end{ruledtabular} \end{table} To understand how movement of the atomic planes in the surface region influences the energy and character of the surface and subsurface states, the B terminated surface electronic structure has been calculated both with and without structural optimization. The relative changes in the (B--Mg) inter-layer separation compared to the bulk values, $\Delta_{ij}$, are provided in Table~\ref{Table1}. We find that the relative change in the inter-layer separation is quite small, in agreement with previous plane wave calculations \cite{Li} using the PBE xc-functional \cite{PBE} where only the top four atomic layers were relaxed. This rigidity of the structure is reflected in the band structure calculations shown in Fig.~\ref{Fig1}. We find that the band structures with and without structural optimization do not differ significantly for the B terminated surface. The energies of the surface and subsurface states at the $\overline{\Gamma}$, $\overline{\text{K}}$and $\overline{\text{M}}$ points for the structurally unoptimized and optimized surfaces are also shown in Fig.~\ref{Fig1}. We find that the B terminated surface has several surface and subsurface states, consisting of three localized bands which are well separated from their corresponding bulk bands. For these localized bands we plot $\varrho_{n,\mathbf{k}}$ in Fig.~\ref{Fig2} at $\overline{\Gamma}$, $\overline{\text{K}}$ and $\overline{\text{M}}$ points. The deepest subsurface state's band (Fig.~\ref{Fig2}(a,b,c)) consists of B $\sigma_1$ orbitals, and is localized mainly in the first and slightly in the second B layer. Specifically, we find for this band $s_{n,\mathbf{k}} \gtrsim 0.66$ throughout the SBZ, as shown in Fig.~\ref{Fig_GPAW_B}. In fact, the B $\sigma_1$ surface band may be described semi-quantitatively by a $+0.6$~eV shift of the top of the B $\sigma_1$ bulk bands. To differentiate between localized states, surface resonances, and bulk states, we recalculated the electronic structure for a thicker slab with six additional MgB$_2$ bulk layers. These results are shown as insets in the upper right corners of the $\varrho_{n,\mathbf{k}}(z)$ plots in Fig.~\ref{Fig2}. For the larger slab we clearly see in Fig.~\ref{Fig2}(b) that the B $\sigma_1$ state is indeed a surface state localized on the topmost B atomic layer. The second group of surface states (Fig.~\ref{Fig2}(d,h,j)) are a combination of Mg $s$ and B $p_z$ orbitals. These $sp_z$ surface states are mainly localized above the surface and between B and Mg layers. Figure~\ref{Fig_GPAW_B} shows that the $sp_z$ surface state is clearly recognizable as it disperses through the SBZ, although some hybridization occurs when crossing the narrow B $\sigma_{2,3}$ band. As with the B $\sigma_1$ band, the $sp_z$ surface band is well described by a $+0.6$~eV shift of the top of the respective bulk band, in this case the $\pi$ bands. We also find for B termination the $sp_z$ states decay slowly into bulk, which is also seen for the extended slab shown as insets in Fig.~\ref{Fig2}(d,h,j). The third group of surface states (Fig.~\ref{Fig2} (g,e,i)) are B $\sigma$ states which are more than $90\%$ localized in the topmost B layer. From Fig.~\ref{Fig_GPAW_B} we see that the B $\sigma_{2,3}$ surface states may be clearly traced throughout the SBZ, at about 0.6~eV above their respective B $\sigma_{2,3}$ bulk bands. Thus, Fig.~\ref{Fig_GPAW_B} shows that for the B terminated MgB$_2$(0001) surface the B $\sigma_1$, $sp_z$, and B $\sigma_{2,3}$ surface bands are near quantitatively described by rigidly shifting the top of the respective bulk bands up in energy by 0.6~eV. \subsection{Mg-terminated MgB$_{\bf 2}$(0001) surface} \begin{figure*} \includegraphics[width=0.7\textwidth]{Fig5} \caption{(Color online) Projected density of surface and subsurface states \(\varrho_{n,\mathbf{k}}(z)\) in the Mg terminated MgB$_2$(0001) surface versus position normal to the surface plane $z$, and Kohn-Sham eigenenergies $\varepsilon_{n,\mathbf{k}}$ relative to the Fermi level $\varepsilon_F$. Results from plane wave PWscf calculations before ({\color{blue}{\bf{---}}}) and after ({\color{red}{\bf{-- --}}}) structural optimization are compared with real-space GPAW calculations for the relaxed structure ($\cdots$). Solid and dotted vertical lines denote positions of the B and Mg atomic layers, respectively. Inserts show $\varrho_{n,\mathbf{k}}(z)$ from plane wave PWscf calculations for a structurally optimized extended supercell model with six bulk unit cells added in the center of the slab.}

\caption{(Color online) Band structure (lines) and energies of the surface and subsurface states shown in Fig.~\ref{Fig4} (symbols) for the Mg terminated MgB$_2$(0001) surface in eV relative to the Fermi level $\varepsilon_F$. Results from plane wave PWscf calculations before ({\color{blue}{\bf{---}}}; {\color{blue}{$\blacksquare$}}) and after ({\color{red}{\bf{-- --}}}; {\color{red}{$\medbullet$}}) structural optimization are compared with real-space GPAW calculations for the relaxed structure ($\cdots$; $\vartriangle$).}

\caption{(Color online) GPAW overlap-based band structure of relaxed Mg terminated MgB$_2$(0001) surface, with \(\psi_{n,\mathbf{k}}\rightarrow\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\) for \(n'\) which maximizes \(\left|\langle\psi_{n,\mathbf{k}}|\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\rangle_{\mathcal{V}}\right|\). Surface states with more than two thirds of their weight in the first vacuum and first two bulk layers, $s_{n,\mathbf{k}} \gtrsim 0.66$, are shown separately ({\color{red}{\bf{---}}}).} \label{Fig_GPAW_Mg} \end{figure} \begin{table} \caption{Change in Mg terminated Mg--B inter-layer separation $\Delta_{i,j}$ between the $i^{\mathrm{th}}$ and $j^{\mathrm{th}}$ layers from the surface, relative to the bulk MgB$_2$ experimental geometry.}\label{Table2} \begin{ruledtabular} \begin{tabular}{lrrrr} &\multicolumn{2}{c}{PWscf LDA\footnote{This Work}}&\multicolumn{2}{c}{VASP PBE\footnote{Ref.~\onlinecite{Li}}} \\\hline $\Delta_{0,1}$[Mg--B]& $-5.2\%$&-9.2 pm& $-3.7\%$&-6.5 pm\\ $\Delta_{1,2}$[B--Mg]& $-0.3\%$&-0.5 pm& $ 1.2\%$& 2.1 pm\\ $\Delta_{2,3}$[Mg--B]& $-1.9\%$&-3.3 pm& $ 0.2\%$& 0.4 pm\\ $\Delta_{3,4}$[B--Mg]& $-0.7\%$&-1.2 pm& $ 0.5\%$& 0.9 pm\\ $\Delta_{4,5}$[Mg--B]& $-0.4\%$&-0.7 pm& $-0.3\%$& -0.5 pm\\ $\Delta_{5,6}$[B--Mg]& $-0.9\%$&-1.6 pm&---& ---\\ $\Delta_{6,7}$[Mg--B]& $-0.75\%$&-1.3 pm&---&---\\ $\Delta_{7,8}$[B--Mg]& $-0.9\%$&-1.6 pm&---&---\\ $\Delta_{8,9}$[Mg--B]& $-0.9\%$&-1.6 pm&---&---\\ $\Delta_{9,10}$[B--Mg]& $-0.9\%$&-1.5 pm&---&---\\ \end{tabular} \end{ruledtabular} \end{table} Even though optimization of the crystal structure only slightly modified the Mg terminated surface (\emph{cf.}~Table~\ref{Table2}) and band structure (\emph{cf.}~Fig.~\ref{Fig3}), modifications of the surface electronic structure are quite radical. Density distributions in Fig.~\ref{Fig4} show that for the unrelaxed surface there exist two types of localized states, an $sp_z$ surface state (Fig.~\ref{Fig4}(a)) and B $\sigma$ subsurface states (Fig.~\ref{Fig4}(d,e)). On the other hand, even though the relaxed surface is only slightly compressed (maximally $-2.7\%$ in the first layer), it induces the appearance of new surface states (Fig.~\ref{Fig4}(b,c)). Such states consist of B $\sigma$ states, which are localized around B layers but penetrate quite deep inside the crystal. At the $\overline{\text{K}}$ point this state may even be considered a surface state resonance. Figure \ref{Fig_GPAW_Mg} clearly shows that these states are more properly associated with the B $\sigma_{2}$ and $\sigma_{3}$ surface states (Fig.~\ref{Fig4} (d,e)). Another consequence of the surface relaxation is that the B $\sigma$ surface states (Fig.~\ref{Fig4}(d,b,e)) are pushed upwards towards the surface B layers. In other words, in the relaxed crystal the $\sigma$ surface states are completely localized in the one or two topmost B layers. As shown in Fig.~\ref{Fig_GPAW_Mg}, the three B $\sigma$ bands have definite surface character throughout the SBZ, where they follow the bottom edge of the B $\sigma$ bulk bands. We also find relaxation lowers the energy of the $\sigma$ subsurface states by approximately $0.1$~eV at the $\overline{\Gamma}$ and $\overline{\text{K}}$ points, and increases the energy of the $sp_z$ surface state from $-1.86$~eV to $-1.78$~eV at the $\overline{\Gamma}$ point, in better agreement with the experimental value of $-1.6$~eV. It should also be noted that the B $\sigma_{3}$ surface band is more localized on the topmost B layer in the real-space GPAW calculation. This is attributable to the inherit difficulties plane wave based methods have in describing state localization, compared to real-space methods. In fact, from Fig.~\ref{Fig_GPAW_Mg} we see that the bottom of all three B $\sigma$ bands have $s_{n\mathbf{k}} \gtrsim$ 0.66 throughout the SBZ. This suggests that adding a layer of Mg shifts the B $\sigma$ surface bands down in energy by about 0.9 eV relative to a the B terminated MgB$_2$(0001) surface. On the other hand, we find that the $sp_z$ surface band is higher up in energy by about 0.9 eV relative to the B terminated surface. In Sec.~\ref{Discuss} we will discuss how this rigid shifting in energy of the surface bands may be understood in terms of charging of the MgB$_2$(0001) surface. \subsection{Li-terminated MgB$_{\bf 2}$(0001) surface} \begin{figure*} \includegraphics[width=0.7\textwidth]{Fig8} \caption{(Color online) Projected density of surface and subsurface states \(\varrho_{n,\mathbf{k}}(z)\) in the Li terminated MgB$_2$(0001) surface versus position normal to the surface plane $z$, and Kohn-Sham eigenenergies $\varepsilon_{n,\mathbf{k}}$ relative to the Fermi level $\varepsilon_F$. Results from plane wave PWscf calculations before ({\color{blue}{\bf{---}}}) and after ({\color{red}{\bf{-- --}}}) structural optimization are compared with real-space GPAW calculations for the relaxed structure ($\cdots$). Solid, dotted, and dashed vertical lines denote positions of the B, Mg, and Li atomic layers, respectively. Inserts show $\varrho_{n,\mathbf{k}}(z)$ from plane wave PWscf calculations for a structurally optimized extended supercell model with six bulk unit cells added in the center of the slab.}

\caption{(Color online) Band structure (lines) and energies of the surface and subsurface states shown in Fig.~\ref{Fig6} (symbols) for the Li terminated MgB$_2$(0001) surface in eV relative to the Fermi level $\varepsilon_F$. Results from plane wave PWscf calculations before ({\color{blue}{\bf{---}}}; {\color{blue}{$\blacksquare$}}) and after ({\color{red}{\bf{-- --}}}; {\color{red}{$\medbullet$}}) structural optimization are compared with real-space GPAW calculations for the relaxed structure ($\cdots$; $\vartriangle$).}

\caption{(Color online) GPAW overlap-based band structure of relaxed Li terminated MgB$_2$(0001) surface, with \(\psi_{n,\mathbf{k}}\rightarrow\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\) for \(n'\) which maximizes \(\left|\langle\psi_{n,\mathbf{k}}|\psi_{n',\mathbf{k}+\Delta\mathbf{k}}\rangle_{\mathcal{V}}\right|\). Surface states with more than two thirds of their weight in the first vacuum and first two bulk layers, $s_{n,\mathbf{k}} \gtrsim 0.66$, are shown separately ({\color{red}{\bf{---}}}).} \label{Fig_GPAW_Li} \end{figure} \begin{table} \caption{Change in Li terminated Mg--B inter-layer separation $\Delta_{i,j}$ between the $i^{\mathrm{th}}$ and $j^{\mathrm{th}}$ layers from the surface, relative to the bulk MgB$_2$ experimental geometry.}\label{Table3} \begin{ruledtabular} \begin{tabular}{lrr} &\multicolumn{2}{c}{PWscf LDA\footnote{This Work}} \\\hline $\Delta_{0,1}$[Li--B]& $-29.8\%$&-52.4 pm\\ $\Delta_{1,2}$[B--Mg]& $-0.6\%$&-1.0 pm\\ $\Delta_{2,3}$[Mg--B]& $-1.2\%$&-2.2 pm\\ $\Delta_{3,4}$[B--Mg]& $-1.5\%$&-2.7 pm\\ $\Delta_{4,5}$[Mg--B]& $-1.2\%$&-2.1 pm\\ $\Delta_{5,6}$[B--Mg]& $-1.5\%$&-2.6 pm\\ $\Delta_{6,7}$[Mg--B]& $-1.4\%$&-2.4 pm\\ $\Delta_{7,8}$[B--Mg]& $-1.9\%$&-3.3 pm\\ $\Delta_{8,9}$[Mg--B]& $-1.5\%$&-2.6 pm\\ $\Delta_{9,10}$[B--Mg]& $-1.5\%$&-2.7 pm\\ \end{tabular} \end{ruledtabular} \end{table} To understand how surface charging, or electro-chemical doping might affect MgB$_2$(0001) surface states, we have directly replaced Mg by Li in the first atomic layer (Mg$\rightarrow$Li). In effect, this approximates the removal of one electron from each of the Mg terminated surfaces. In other words, replacing Mg by Li resembles a positively charged Mg terminated (Mg$^{+1}$) surface. This allows us to directly probe the effect of surface charging of the Mg terminated surface without the need for compensating background charges in the calculation. We have then performed structural optimization of the Li terminated MgB$_2$(0001) surface to model how Li termination changes the surface states. From Table~\ref{Table3} we see that the smaller atomic radius of Li induces a substantial relaxation, with the relaxed Li atomic layer moved down by 52.4 pm towards the B atomic layer. Both the surface charging and the crystal structure modification strongly change the electronic structure of the surface states. Figure \ref{Fig5} shows that the lowest B $\sigma_1$ surface resonance band for the structurally unoptimized surface (Mg$\rightarrow$Li$\sim$Mg$^{+1}$), is slightly separated from the upper edge of the bulk B $\sigma_1$ band. The density distributions of such states are shown in Fig.~\ref{Fig6}(a,b,c,f), where we see that B $\sigma_1$ surface resonances are mainly localized in the four topmost B layers. Structural optimization causes a slight downward shift of the surface resonance B $\sigma_1$ bands and induces a strong interaction/hybridization with the bulk B $\sigma_1$ band, so that they decay into the bulk states. I.e., it increases the amplitude of the propagation of the resonance state into the bulk, as can be seen at the $\overline{\Gamma}$ point in Fig.~\ref{Fig6}(a). For $\overline{\text{K}}$ and $\overline{\text{M}}$ points, the B $\sigma_1$ surface resonances cannot even be distinguished from the normal bulk states and are not shown. This is seen clearly in Fig.~\ref{Fig_GPAW_Li}, with the relaxed Li terminated B $\sigma_1$ surface states becoming increasingly bulk-like between $\overline{\Gamma}$ and K, until $s_{n,\mathbf{k}} < 0.66$. Relaxation also causes a $0.4$~eV reduction of the binding energy of the $sp_z$ supra-surface state band. This is expected because relaxation causes contraction of the topmost Li-B interstitial region (\emph{cf.}~Table \ref{Table3}) where the $sp_z$ surface state is mostly located (\emph{cf.}~Fig.~\ref{Fig6}). The binding energy of the $sp_z$ surface state at $\overline{\Gamma}$ point is strongly reduced, from about $2.6$~eV to $0.1$~eV, comparing with the B terminated surface. This suggests that the binding energy of the $sp_z$ surface state at the $\overline{\Gamma}$ point may be strongly influenced by surface contamination or doping of the B terminated surface, as we will discuss in Sec.~\ref{Discuss}. We also find the more localized B $\sigma_2$ and $\sigma_3$ bands are strongly modified by Li termination. As can be seen in Fig.~\ref{Fig5} for Mg$\rightarrow$Li$\sim$Mg$^{+1}$, the B $\sigma_{2}$ and $\sigma_3$ surface bands are always slightly separated from the upper edge of the bulk B $\sigma_{2}$ and $\sigma_3$ bands, with their density highly localized in the first one or two B layers, as seen in Fig.~\ref{Fig6}(g,e,h). These states resemble the highly localized B $\sigma_{2}$ and $\sigma_3$ surface states in the B terminated surface. However, surface relaxation shifts the localized B $\sigma_2$ and $\sigma_3$ bands down in energy. This causes a strong interaction between localized and bulk $\sigma$ states, with the surface state decaying into the bulk states. This is clearly illustrated by the densities of these states for the relaxed Li terminated structure shown in Fig.~\ref{Fig6}(e,h). Here we see that the $\sigma$ localized state now penetrates substantially into the crystal. By plotting the same distributions for an extended crystal structure (\emph{cf.}~insets of Fig.~\ref{Fig6}(e,h)) we see that these states are actually localized state resonances or even pure bulk states at the $\overline{\text{K}}$ point. From Fig.~\ref{Fig_GPAW_Li} we see that the B $\sigma_{2}$ and $\sigma_3$ surface states strongly hybridize with the bulk $\sigma$ states, with $s_{n,\mathbf{k}} < 0.66$ over much of the SBZ. From this we make two observations. First, removing charge from the Mg terminated surface would strongly affect the surface and subsurface states. Second, for the relaxed Li terminated surface all subsurface states hybridize into localized state resonances or bulk states, while the $sp_z$ surface state survives. We will discuss how these affects may be understood and controlled, with the goal of tuning the surface states of MgB$_2$, in the following section. \section{Discussion} \label{Discuss} \begin{figure} \includegraphics[width=0.85\columnwidth]{Fig11} \caption{Surface and subsurface state energies at the $\overline{\Gamma}$ point, \(\varepsilon_{n,\overline{\Gamma}}\) versus surface termination and doping of the MgB$_2$(0001) surface by B, Mg$\rightarrow$Li$\sim$Mg$^{+1}$, Li, and Mg, in eV relative to the Fermi level $\varepsilon_F$, for (a) B $\sigma_{2}$ and $\sigma_3$, (b) $sp_z$, and (c) B $\sigma_1$ states. Results from plane wave PWscf calculations before ({\color{blue}{$\blacksquare$}}) and after ({\color{red}{$\medbullet$}}) structural optimization are compared with real-space GPAW calculations for the relaxed structures ($\vartriangle$). The B $\sigma_{2,3}$ and B $\sigma_1$ bulk band regions are shaded gray. Fits to the surface state positions ({\bf{-- --}}) are provided as guides to the eye.}

36 — 1105.2810

\caption{(Color) (a): the crystallographic structure of \Livcu \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$^{+}$). (b): the main in-- and inter-chain exchange paths, $J_1$, $J_2$, and $J_3$,$J_4$, $J_5$ marked by solid red arcs, broken red line, blue and green arcs , respectively. }

\caption{(Color) Magnetization vs. applied magnetic field. Experiment (\textcolor{green}{$\circ $} from %\textcolor{green}{ \bullet } ) are taken from Ref.\\cite{Svistov11}), theory: blue line - set of Ref.\\cite{Enderle10}, red line - our set \cite{Drechsler11} and black line- the $U=5$~eV- set from Ref.\\cite{Koo11} and $g=2.3$ for all sets. The DMRG calculations were performed for $L=512$ sites at $T=0$. Inset: 'entire' field range. }

37 — 1105.3076

\caption{{\cred{ (Color online) 3D map of the local bases $\{ \hat{\bx}_a, \ x_1, \ x_2, \ z \}$ introduced at selected nodes of a triangulated model of the RBC membrane ($x_1$: yellow, $x_2$: blue, $z$: cyan) }}.}

\caption{ {\cred{ (Color online) 3D density plots of the LME approximations ${H}_N$ to the mean curvature $H$ of of a CGMD model of the RBC membrane %CDL why not just use reference numbers \cite{Marcelli:2005,Hale:2009} at different simulation times $t$, for $\bar{\beta}=125$ and $m=10$ ($ {\hat {\bar H}}_N = {\bar H}_N \times 10^{5} \ \mbox{\AA}^{-1}$; lower bound of the color bar: $-11.5 \times 10^{-5} \ \mbox{\AA}^{-1}$; upper bound: $+2.5 \times 10^{-5} \ \mbox{\AA}^{-1}$). }}}

\caption{ {\cred{ (Color online) 3D density plots of the LME approximations ${K}_N$ to the Gaussian curvature $K$ of of a CGMD model of the RBC membrane \cite{Marcelli:2005,Hale:2009} at different simulation times $t$, for $\bar{\beta}=125$ and $m=10$ (lower bound of the color bar: $-2 \times 10^{-9} \ \mbox{\AA}^{-2}$; upper bound: $+7 \times 10^{-9} \ \mbox{\AA}^{-2}$). }}}

\caption{ {\cred{LME estimates of the zero-temperature bending rigidity of the RBC membrane model presented in Section \ref{RBC}, at different times of a MD simulation.}}}

38 — 1105.3173

\caption{(color online) IT-NCSM ground-state energies for \elem{He}{4} and \elem{Li}{6} as function of $N_{\max}$ for the three types of Hamiltonians (see column headings) for a range of flow parameters: $\alpha=0.04\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.05\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.0625\,\text{fm}^4$ (\symboltriangle[FGGreen]), $0.08\,\text{fm}^4$ (\symbolbox[FGViolet]), and $0.16\,\text{fm}^4$ (\symbolstar[FGLightBlue]). Error bars indicate the uncertainties of the threshold extrapolations. The bars at the right-hand-side of each panel indicate the results of exponential extrapolations of the individual $N_{\max}$-sequences (see text).}

\caption{(color online) $N_{\max}$-extrapolated ground-state energies of \elem{He}{4} and \elem{O}{16} as function of the flow parameter $\alpha$ for the NN-only (\symbolcircle[FGBlue]), the NN+3N-induced (\symboldiamond[FGRed]), and the NN+3N-full Hamiltonian (\symboltriangle[FGGreen]).}

39 — 1105.3524

\caption{\lya{} line shapes for a $10^{11}\,\msolar{}$ galaxy at $z=2.2$ in various environments. The dotted curve is the original \lya{} line shape and the grey dashed curve is the transmitted line shape assuming a mean \igm{} density and UV background. The solid \blackbold{} (\redbold{}) curves show the impact of an increased \igm{} density and UV flux from the nearby quasar with a host halo size of $2\times10^{13}\,\msolar{}$ ($10^{12}\,\msolar{}$). The dash-dotted curves include the increased \igm{} density but without the quasar's contribution to the incident UV flux.}

\caption{Averaged \lya{} transmission as a function of radius from a central quasar with host halo size (\blackbold{} curves) $2\times10^{13}\,\msolar{}$ and (\redbold{} curves) $10^{12}\,\msolar{}$. The grey dashed curve is the average transmission assuming a mean \igm{} density and UV flux. The solid curves include the density and UV flux from the nearby quasar and the dash-dotted curves include the increased density but with no UV boost.}

\caption{Predicted number of galaxies contained within the Francis--Bland-Hawthorn volume as a function of modeled duty cycle for a quasar with host halo size (\blackbold{} curves) $2\times10^{13}\,\msolar{}$ and (\redbold{} curves) $10^{12}\,\msolar{}$. The grey dashed curve is the number of galaxies expected in the field. The long-dashed lines show the enclosed count assuming a constant $\igmtrans{}=0.5$. The solid (dash-dotted) curves include clustering and transmission effects including (excluding) the boosted UV flux from the quasar. The upper and lower dotted curves for each quasar show the enclosed count excluding galaxies with $\virialtemp{} < 5\times10^{5}\,\kelvin{}$ ($\lesssim6.6\times10^{10}\,\msolar{}$) and $\virialtemp{} < 10^{6}\,\kelvin{}$ ($\lesssim1.9\times10^{11}\,\msolar{}$) respectively.}

40 — 1105.3768

\caption{ For case \#\,4. (a): evolution of the electron spectrum in one of the blob zones. (b): SEDs. In both panels, the time sequence is: black, red, blue, cyan. Times are in the observer's frame. Model parameters:$B=0.1$\,G,$\delta=33$, sizes: $z=1.33\times10^{16}$\,cm,$r=10^{16}$\,cm,$n_e=0.4$\,cm$^{-3}$, injection rate $q=3.17\times10^{39}$\,erg/s without shock increasing by a factor of 8 with the shock. Time-scales: acceleration$t_{\rm acc}=z/c$, and achromatic loss $t_{\rm loss}=z/c$. The shock begins to cross the blob at $t=60$\,ks.\label{fig:case4_electrons_and_seds} }

\caption{ Case \#4. (a): Light curves in three energy bands. (b):\gray vs. \xray flux correlation. Colors represent different times: black ($<100ks$), red ($100-120$ ks), blue ($120-140$ ks), cyan ($140-160$ ks), magenta ($160-180$ ks), yellow ($>180$ ks). \label{fig:case4_lc_and_ff} }

41 — 1105.4582

\caption[Selected stuff]{A subset of data comparing scores on the Big Five personality traits and naturalness as given by native speakers of American English (left half of the page) and Arabic (right half of the page). Blue, white, and pink bars correspond to negative, neutral, and positive valences of the linguistic features respectively. Dialogue acts listed along the horizontal axis are a greeting, question-answer pair, disagreement, and apology. Error bars the 95\% confidence intervals, brackets above the plots correspond to p-values of paired t-tests at significance levels of 0.05 (\textasteriskcentered) and 0.01 (\textasteriskcentered\textasteriskcentered) after Bonferroni correction.}

42 — 1105.5699

\caption{ Artifacts at the L24 band. The target object is IRAS 22396$-$4708. The identified artifacts (A, B, and C) are indicated. The image is displayed in a logarithmic scale. The entire field-of-view (256$\times$256 pixels) is shown. The black regions in the center are saturated pixels. The structures seen in the right edge are the slit mask. \textcolor{red}{The diffraction spikes and banding (see \S~\ref{psf}) can also be seen in the figure and are overlapped on the artifacts, which can be separated from the artifacts with the separation procedure (see \S~\ref{sub_Lartifact}) and are finally included in the PSF.}\label{fig00}}

\caption{ \small Example of the image reconstruction process of the IRC mid-infrared data investigated in the present study (M81: pointing ID 5020062.1 and 5020063.1). (a) Artifact-subtracted image of S11, (b) S11 image reconstructed with equation (\ref{eqconv}), (c) artifact-subtracted image of L24, (d) L24 image reconstructed with equation (\ref{eqconv}), (e) the L24/S11 color map produced from the L24 image (c) and the S11 image (a) convolved with a simple Gaussian to adjust the FWHM to 6\arcsec.7 (green line in Figure~\ref{radconv}), and (f) the L24/S11 color map produced from the reconstructed images (b) and (d). Arrows indicate bright \ion{H}{2} regions and the galactic center of M81, where ring-like artifacts appear in (c) and (e). After the image reconstruction procedure, the ring-like artifacts seen in (c) and (e) disappear in (d) and (f). \textcolor{red}{The blue line in region A overlaid on (c) and (d) indicates the position of the surface brightness profiles shown in Figure~\ref{fig15} and the green boxes indicate the regions used to estimate the general fluctuation level (see \S~3).} \label{rec1}}

\caption{\textcolor{red}{Surface brightness profiles at L24 along the line shown in Figures~\ref{rec1}c and d. The dashed line represents the profile before image reconstruction (Figure~\ref{rec1}c); the solid line shows that after image reconstruction (Figure~\ref{rec1}d). The pixel scale is 2\arcsec.38, which corresponds to about $42$ pc assuming the distance of 3.6 Mpc to M81. A bright HII region is located at $\Delta Y = 12.5$ pixels and the ring-like pattern appears at $\Delta Y \sim 7.5$ pixels and $\sim 17.5$ pixels.} \label{fig15}}

43 — 1105.5821

\caption{{\bf Mismatch feature space representation.} The mismatch feature space representation of a segment of a protein sequence ...\red{AQGPR}IYDDT\cyan{CQHPS}WWMNFE\darkgreen{YRGSP}...}

44 — 1106.1981

\caption{Density profiles of impurity ions against minor radius, both without (a) and with (b) a sheared radial electric field. Simulated ions are {H}e$^{2+}$ (\full), {C}$^{6+}$ (\broken), {N}e$^{10+}$ (\chain) and {W}$^{20+}$ (\dotted). Vertical lines in (b) indicate the peak strength and full-width half-maximum positions of the sheared radial electric field.}

45 — 1106.2231

\caption{(a) Initial state and (b) state after evolution by the Duffing system of Section~\ref{sec:sys}. The \protect\coloronline{blue}{dark} region is surrounded by a transport barrier, while the \protect\coloronline{green}{light} region is not. Particle positions are shown as \protect\coloronline{red dots}{dots}.}

\caption{Simplified loop diagrams of the loops which do not show steady growth. (a) Loop containing the limit-cycle trajectories (\protect\coloronline{red}{dark} punctures --- 3, 5, and 7). (b) Loop containing the inner chaotic trajectories (\protect\coloronline{pink}{light} punctures --- 1, 2, 4, and 6). (c) Double-loop consisting of the (a) and (b) loops.}

\caption{Poincar\'e section (stroboscopic map) for a vat of viscous fluid stirred by a rod (see inset for the rod's path). A chaotic sea and several islands are visible, as well as regular orbits near the wall. The squares indicate the initial position of the~$40$ trajectories, with non-growing loops detected by the pair-loop method drawn in.}

46 — 1106.4224

\caption{The 1-month E$> 100$ MeV light curve (2-day time-bin) of PKS 1830-211 (2010 October 8 - 2010 November 8). The two horizontal lines represent the 1-month and the 4-day \gray enhancements. On the right $y$-axis we report the observed \gray SED. The downward triangles indicate the 2-$\sigma$ \gray upper limits. The plot insert shows the soft X-ray light curve.}

\caption{The observed SED of PKS 1830-211 for the \gray flare (red filled circles), for the 1-month enhancement (blue filled circles) and for the average state (black points). In detail, the black open circles are the non-simultaneous data from radio to hard X-rays (Pramesh Rao \& Subrahmanyan 1998, Courbin et al. 2002, 7-year INTEGRAL survey, respectively). We report as open stars Planck data taken from the ERCSC of the first all-sky survey of Planck. The red downward arrows represent the near-IR{\it SMARTS} and the INTEGRAL/IBIS (20-40 keV) upper limits. Solid lines represent the SED models for the three \gray states (black for the average (EGRET), blue for the 1-month, red for the flare). The models were magnified by a factor of ten for the lensing.}

\caption{Model Parameters for 2010 \gray activity of 1830.}

47 — 1106.4482

\caption{\label{fig:Wu-MDR} Velocity statistics of the transitional boundary layer simulation of \cite{wu2009direct} for two states whose velocity profiles approach closely Virk's asymptotic velocity profile (Eq.~\ref{eq:virk}). (a) Shows \eqnref{eq:virk} \color{cyan}$\boldsolid$\color{black}, $Re_\theta=200$ \solid, $Re_\theta=300$ \dashed, and $Re_\theta=900$ \dotted (fully turbulent flow). (b) Compares Reynolds shear stress, $-\avg{uv}^+$, for the three flow states as shown in (a). (c) Plots the distribution of the skin friction coefficient $C_f$ as a function of the Reynolds number based on the local momentum thickness $\theta$. Arrows point to the statistical profiles plotted in (a) and (b). }

\caption{\label{fig:DR}Polymeric flows at MDR. (a) Shows MDR flow time and space-averaged mean velocity profile \solid and fluctuations of time-dependent, space-averaged mean velocity (error bars), \eqnref{eq:virk} $\color{cyan}\boldsolid$\color{black}, $Re_\theta=200$ \dashed, and $Re_\theta=300$ \dotted. (b) Compares Reynolds shear stress, for the three flow states as shown in (a). (c) Plots the time-evolution of pressure gradient for different flow and initial conditions: \dashed Newtonian initiated with isotropic turbulent entrance flow (turbulent intensity $u'/U_b=0.01$), \solid MDR flow ($L=200,We_\tau=720$) initiated with viscoelastic isotropic turbulent entrance flow ($u'/U_b=0.005$), and \dotted MDR flow ($L=160,We_\tau=200$) initiated from a fully developed turbulent channel flow with $\ten{C}=\ten{I}$ condition for the polymer field. }

48 — 1106.4669

\caption{ The results of more theoretical methods not mentioned in Fig. 3. Notations are detailed in the main text. (a)Absorption wavelengths calculated using TD-CAM-B3LYP({\color{red}$\bullet$})~\cite{R-CAM}, TD-$\omega$B97X-D({\color{blue}$\bullet$})~\cite{R-wB97XD}, TD-$\omega$B97({\color{magenta}$\bullet$})~\cite{R-wB97}, TD-LC-$\omega$PBE({\color{darkyellow}$\bullet$})~\cite{R-LCwPBE} (LC hybrid functionals), TD-PBE1PBE({\color{wine}$\blacktriangle$})~\cite{R-PBE1PBE} (a global hybrid functional) and CIS({\color{olive}$\star$})~\cite{R-CIS} (TD-HF with Tamm-Dancoff approximation) vs. absorption wavelengths from EOM-CCSD($\diagup$)~\cite{R-EOM}. Data of each method for the same excitation lies on the same vertical line. (b) Oscillator Strength: line denotes the results from EOM-CCSD calculation. (c) $\sigma_{M}$ of the Type-A$\sim$Type-G excitations. }

49 — 1106.4963

\caption{(a) Average displacement field for a time lag corresponding to an intruder displacement over its radius $R = 10$ mm and for a coarse-graining size $\Delta W=d_s$. The color code represents the amplitude of the coarse-grained displacement normalized by the intruder radius and the vectors indicate the direction of grain motion. Note the horizontal distance $l_v$ between the intruder center and a vortex center. (b) Horizontal positions of left and right vortex centers as a function of time ( density contrast $\frac{\Delta\rho}{\rho}$=16.15), ((\textcolor{red}{$\bullet$}) right vortex and (\textcolor{blue}{$\bullet$}) left vortex). (c) Mean horizontal distances $\bar{l}_v$ between the intruder and the vortex centers as a function of the radius $R$ (symbol colors as in (b)).}

50 — 1106.4966

\caption{\label{fig:Heisenberg2D} Lower energy bounds and correlators for the 2D XXZ model, complemented by QMC data calculated for a square lattice of $16\times 16$ sites with periodic boundary conditions, and inverse temperature $\beta=96$. The QMC energies for $J_z=0,1$ coincide with earlier results (\blue{$\blacklozenge$}) from Refs.~\cite{Sandvik1997-56,Sandvik1999-60}. The QMC error bars would be smaller than the line width, and another simulation with a $32\times 32$ lattice produced visually indistinguishable results. As SDP constraint subsystems $\Omega_k$ [Eq.~\eqref{eq:constraintOp}] we chose $L_k\times L_k$ squares with $K=4$ and $L_1,L_2,L_3,L_4$ as specified in the legend.}

51 — 1106.5024

\caption{Intrinsic field strength $\B$, comparing the original model magnetogram to the bin-by-30 results, for different binning approaches. (a) original {\it vs.} {\bf ``instrument''}, (b) original {\it vs.} {\color{green} ``bicubic''}, (c) original {\it vs.} {\color{red}``average''}, and (d) original {\it vs.} {\color{violet} ``sampled''}. For all, the $x=y$ line is also plotted for reference, and on the $x$-axis (``Original'') are plotted all the values represented by the single resulting bin-30 pixel in question, whose value is plotted on the $y$-axis. Every other point in the binned magnetogram is shown, and every 3rd point of the 900 underlying values is plotted. The colors for these plots will be used consistently below.}

\caption{\small Average intrinsic field strength $\frac{1}{N} \sum{\B}$ as a function of binning factor (top x-axis), for the four binning methods: {\bf ``instrument'' ($\Diamond$)}, {\color{green}``bicubic'' ($\Box$)}, {\color{red} ``average'' ($\times$)}, and {\color{violet} ``sampled'' ($\triangle$}). The three panels show, respectively, the full magnetogram, an ``umbral'' area and a ``plage'' area (see Figure~\ref{fig:flowers}). For each binning, $N$ varies but the same sub-area of the ``Sun'' is covered; when non-integer pixel numbers result, that bin factor is omitted. The original model field is sampled at an arbitrarily-set $0.03\arcsec$, the resulting ``pixel sizes'' are indicated (bottom x-axis). For these and all similar plots (except where noted), the y-axis ranges are kept consistent between the target areas for direct comparisons. Here, the effects are minimal for the full magnetogram and the ``umbra'', but have a much larger magnitude and differ between the binning methods in the ``plage'' area.}

\caption{\small Cumulative probability distributions, comparing that for the full-resolution synthetic map to the bin-30 results, for the three fields of view (entirety, ``umbra'', and ``plage'' areas). For each, CPD curves are plotted for the {\color{blue} original resolution}, the {\bf instrument} method, and the {\color{green} bicubic}, {\color{red} average}, {\color{violet} sampled} post-facto approaches. Top: For the intrinsic field strength $\B$; Bottom: For the vertical electric current density $\Jz$.}

\caption{\small Again for the three fields of view, summaries of the Kolmogorov-Smirnoff tests as a function of binning factor, for the field strength $\B$. {\bf Top:} the ``D'' statistic and {\bf Bottom:} the probability ``P'' that the two samples considered are {\it different} (see text). Shown are four curves, original resolution {\it vs.} {\bf ``instrument'' ($\Diamond$)}, {\color{green} ``bicubic'' ($\Box$)}, {\color{red} ``average'' ($\times$)}, and {\color{violet} ``sampled'' ($\triangle$}) magnetograms.}

\caption{\small Following Figure~\ref{fig:flowers_BB}, the the average field strength over the target area, $\frac{1}{N} \sum{\B}$ as a function of binning factor (top $x$-axis), for the four binning methods ({\bf ``instrument'' $\Diamond$}, {\color{green}``bicubic'' $\Box$}, {\color{red} ``average'' $\times$} and {\color{violet} ``sampled'' $\triangle$}), focusing on three areas as indicated: the full magnetogram, an ``umbral'' area and a ``plage'' area, as depicted in Figure~\ref{fig:hinode}. With the original {\it Hinode}/SP scan resolution of $0.15\arcsec$, the resulting pixel sizes are also indicated (bottom $x$-axis). }

\caption{\small Cumulative probability distributions, for the full-resolution data and the bin-10 results, for the three fields of view (entirety, ``umbra'', and ``plage'' areas). For each, CPD curves are plotted for: {\color{blue} original resolution}, {\bf instrument-binning}, {\color{green} bicubic}, {\color{red} average} and {\color{violet} sampled} approaches. Top: intrinsic field strength $\B$; Bottom: vertical electric current density $\Jz$.}

\caption{\small Summaries of the Kolmogorov-Smirnoff tests as a function of binning factor, for the intrinsic field strength $\B$ over the three fields of view. Top: the ``D'' statistic; Bottom: the probability that the two samples are {\it different} (see text). Shown are curves for the original resolution {\it vs.} the ({\bf ``instrument'' $\Diamond$}, {\color{green}``bicubic'' $\Box$}, {\color{red} ``average'' $\times$} and {\color{violet} ``sampled'' $\triangle$}) vertical current distributions.}

52 — 1106.5652

\caption{Left panel: rescaled values $\tilde{\lambda}^{(r)}$, defined as $\tilde{\lambda}_A^{(r)}=\tilde{\lambda}_A$ and $\tilde{\lambda}_{B,C}^{(r)}=\tilde{\lambda}_{B,C}/\sqrt{2}$, plotted as a function of $\overline{\rho}$, showing a collapse of the data (same data as in the right panel of Fig.~\ref{fig-variance}). Right panel: differences in the locally-measured chemical potential of system $B$ and $A$ ({\color{red} $\times$}), and system $C$ and $A$ ({\color{blue} +}) as a function of $\overline{\rho}$ compared to the analytical prediction $\alpha(\overline{\rho})=1/(3\overline{\rho}$ (black line) as discussed in Sect.~\ref{consistency}.}

53 — 1106.6042

\caption{{\em Left:} The central temperature of the Cold Spot as a function of the radius $r_{||, {\rm emission}}$, chosen parallel to the line of sight, at which the surface of last scattering intersects the sphere. In solid red we show the temperature fluctuation as calculated by explicit solving of the geodesic equations, which takes into account both the Bardeen potential (ordinary Sachs Wolfe effect, dashed green curve) and the velocity along the line of sight (the dashed-dotted blue curve, which is obtained by taking the difference between the solid red and dashed green curves). {\color{mydarkdarkred} The velocity is automatically taken into account since the coordinates are comoving with the matter, and we assume emission from a fixed coordinate. } The Bardeen-potential curve is identical to the curve shown in Fig~\ref{fig:tempprofile}, the difference being that in that figure the Bardeen potential is the only contribution to the temperature fluctuation, since the surface of last scattering there intersects the sphere at $r_{||, {\rm emission}}=0$. The fact that in this figure the difference curve (blue dashed dotted) is anti-symmetric in $r_{||, {\rm emission}}=0$, provides a good cross-check between the different calculations of the temperature of the Spot. {\em Right:} A schematic presentation of three different radii at which the last scattering surface (the three upright planes) intersect the spherical object. The cone on the right of the illustration represents the observer. }

54 — 1107.2151

\caption{\label{Figure:CrossSections}(Color online) Differential cross sections for $\gamma p\to p\pi^0$ from CBELSA/TAPS ({\large\color{red}$\bullet$}) plotted versus cos\,$\theta_{\rm c.m.}$. For comparison, CB-ELSA data~\cite{Bartholomy:2004uz} are represented by ({\tiny\color{blue}$\blacksquare$}) and CLAS data~\cite{Dugger:2007bt} by ({\large\color{cyan}$\circ$}). Note that the CLAS cross sections were extracted for 50-MeV-wide energy bins; for $E_\gamma < 1.3$~GeV, the same CLAS results are thus shown in the corresponding 25-MeV-wide bins chosen for this analysis. The solid blue line shows the solution of the Bonn-Gatchina partial-wave analysis~\cite{Anisovich:2010an}, and the dashed black line represents the current SAID solution (SP09)~\cite{Dugger:2009pn}. The data points include statistical errors only; the total systematic error is given as error bands at the bottom of each plot.}

\caption{\label{Figure:CrossSectionsDeg}(Color online) Differential cross sections for $\gamma p\to p\pi^0$ from CBELSA/TAPS ({\large\color{red}$\bullet$}) plotted versus $\theta_{\rm c.m.}$. For comparison, CB-ELSA data~\cite{Bartholomy:2004uz} are represented by ({\tiny\color{blue}$\blacksquare$}) and CLAS data~\cite{Dugger:2007bt} by ({\large\color{cyan}$\circ$}). Note that the CLAS cross sections were extracted for 50-MeV-wide energy bins; for $E_\gamma < 1.3$~GeV, the same CLAS results are thus shown in the corresponding 25-MeV-wide bins chosen for this analysis. The solid blue line shows the solution of the Bonn-Gatchina partial-wave analysis~\cite{Anisovich:2010an}, and the dashed black line represents the current SAID solution (SP09) ~\cite{Dugger:2009pn}. The data points include statistical errors only; the total systematic error is given as error bands at the bottom of each plot.}

\caption{\label{Figure:Selected}(Color online) The differential cross sections for $\gamma p\to p\pi^0$ at $E_\gamma\in [1350,1400]$~MeV (left), $E_\gamma\in [1750,1800]$~MeV (center), and $E_\gamma\in [1900,1950]$~MeV (right). The angle shown is the pion center-of-mass scattering angle. The experimental data are from the CBELSA/TAPS measurements presented here (filled circles {\large\color{red}$\bullet$}), CB-ELSA~\cite{Bartholomy:2004uz}~(open stars {\color{blue}\ding{73}}), GRAAL (left)~\cite{Bartalini:2005wx}~(open boxes {\scriptsize$\square$}), SPring-8~(center and right)~\cite{Sumihama:2007qa}~(open boxes {\scriptsize$\square$}), CLAS~\cite{Dugger:2007bt}~(open circles {\large\color{cyan}$\circ$}), Althoff et al.~(left) \cite{Althoff:1979mc}~(open triangles {\color{green}$\vartriangle$}), and Brefeld et al.~(right) \cite{Brefeld:1975dv}~(open triangles {\color{green}$\vartriangle$}). The dashed lines represent the current SAID solution (SP09)~\cite{Dugger:2009pn}, and the solid lines represent the latest solution of the Bonn-Gatchina PWA~\cite{Anisovich:2010an}, which includes the new data presented here. The total systematic error is given again as an error band at the bottom of each figure.}

\caption{\label{Figure:Comparison}(Color online) Fixed-angle excitation functions for the reaction $\gamma p\to p\pi^0$; the angle shown is the pion center-of-mass scattering angle. The experimental data are from the new analysis presented here (filled circles {\large\color{red}$\bullet$}), CB-ELSA~\cite{Bartholomy:2004uz} ({\tiny\color{blue}$\blacksquare$}), CLAS~\cite{Dugger:2007bt}~(open circles {\large\color{cyan}$\circ$}), and LEPS~\cite{Sumihama:2007qa} (filled {\large $\star$}). Note that the CLAS cross sections were extracted for 50-MeV-wide energy bins; for $E_\gamma < 1.3$~GeV, the same CLAS results are shown in the corresponding 25-MeV-wide bins chosen for this analysis.}

55 — 1107.3464

\caption{\label{fig:tcal} (Color online) Caloric temperature curves $T(e)=\beta^{-1}(e)$ for a selection of elastic, flexible polymers with chain lengths in the interval $N=13,\ldots,309$ (from right to left). Curves for chains with magic length ($N=13,55,147,309$) are bold. The relevant inflection points, indicating the conformational transitions on the basis of our analysis method, are marked by \textcolor{blue}{$\bullet$} symbols. }

\caption{\label{fig:tn} (Color online) Transition temperatures $T_{\rm tr}(N)$ of conformational transitions for small elastic polymers with chain lengths $N=13,\ldots,309$ in the liquid-solid and solid-solid transition regimes, obtained from inflection-point analysis. First-order transition points are marked by red symbols (\textcolor{red}{$\bullet$}), second-order transition points by blue symbols (\textcolor{blue}{$\times$}). Also shown is a fit for the liquid-solid transition temperature towards the thermodynamic limit $N\to \infty$ (dashed line). }

56 — 1107.5700

\caption{\small \textbf{Vortex core resonant reversal by a single microwave pulse.} \textbf{a}, Efficiency of the $\Pi_-$ and $\Pi_+$ pulses (see panel \textbf{b}) to reverse the vortex core at $\mu_0 H=65$~mT measured as a function of power and frequency (stepped by increments of 0.3~dBm and 1.2~MHz). For each pulse type, a coloured pixel ($\Pi_-$: red shade, $\Pi_+$: blue shade) marks a successful reversal. The transparency gives the switching probability averaged over 16 attempts. Experiments corresponding to three values of the pulse duration $w$ are displayed. For each $w$ we define ($f_+^*$,$P^*$), the optimal working point of the $\Pi_+$ pulse located at the bottom of the corresponding contour plot (see \color{red}$\blacksquare$\color{black}). \textbf{c}, Numerical calculation (see Methods) of the experiments presented in \textbf{a}. This calculation is not valid within the shaded area, where multiple vortex core reversals can occur. \textbf{d, e}, Experimental (\color{red}$\blacksquare$\color{black}) and calculated (lines) dependences on $w$ of the optimal frequency $f_+^*$ (\textbf{d}) and of the optimal pulse energy $E^*=P^*w$ (\textbf{e}). The experimental points are obtained from the analysis of data sets similar to those presented in \textbf{a}, where $w$ is varied from 1~$\mu$s down to 3~ns (the three \color{red}$\blacksquare$\color{black}~close to the minimum energy in \textbf{e} are inferred from those displayed in \textbf{a} ). The best agreement is obtained for a damping ratio $d_\text{forced}^*=0.018$, i.e., a characteristic decay time $\tau_\text{forced}=35\pm4$~ns. The absolute value of the energy is also fitted in the calculation, which enables to extract the critical velocity for vortex core reversal, $V_c\simeq 190$~m/s (see Methods). Error bars on $f_+^*$ and $E^*$ are absolute minima and maxima resulting from experimental uncertainties in \textbf{a} on the optimal working point ($f_+^*$, $P^*$) associated to each pulse duration.}

\caption{\small \textbf{Phase coherent control of vortex core reversal.} \textbf{a}, Numerical calculation (see Methods) of the double pulse sequences presented in Fig.3. The best agreement is obtained for a characteristic decay time $\tau_\text{free}=53\pm6$~ns in the free regime. \textbf{b}, Associated vortex core trajectory (left) and velocity (right) vs. time plotted for two $\Pi_+$ pulses with settings $\tau=60$~ns and $f=253$~MHz (see pixel \color{yellow}$\blacksquare$\color{black}~in \textbf{a}). For these settings, the vortex core is reversed when the phase difference between the pulses is equal to zero (top, see star) and not reversed when it is equal to $\pi$ (bottom).}

57 — 1107.5930

\caption{\blue[Several graphical representations for the classifier] with probability estimates (0.95, 0.90, 0.90, 0.85, 0.70, 0.70, 0.70, 0.55, 0.45, 0.20, 0.20, 0.18, 0.16, 0.15, 0.05) %(0.05 0.15 0.16 0.18 0.20 0.20 0.45 0.55 0.70 0.70 0.70 0.85 0.90 0.90 0.95) and classes (1, 0, 1, 0, 0, 1, 0, 0, 0, 0, 0, 0, 0, 1, 0) %(0 1 0 0 0 0 0 0 0 1 0 0 1 0 1). \blue[Left: ROC curve] (solid) \blue[and convex hull] (dashed). Middle: cost lines and optimal cost curve against cost proportions. Right: cost lines and optimal cost curve against skews.}

58 — 1108.0659

\caption{\label{Movie} The video has been uploaded to the Data Conservancy Pilot Project of arXiv repositories. Also, this and more videos are available to download in: \textcolor{blue}{ \url{http://www.unizar.es/departamentos/fisica_mat_condensada/people/hsruizr/} } \\ \textit{Movie Caption:} \\ Sketch of one of the analized cases along this letter. In this case, $B_{y}^{max}=2$, and $J_{tr}^{peak}=0.75 J_{c}$ (Figure~ \ref{Fig_1} is a frame of this video). Units are $(\mu_{0}/4\pi)J_{c}R$ for B, $\pi R^{2}J_{c}$ for $I_{tr}$, $J_{c}R^{3}$ for M, and $(\mu_{0}/4\pi)J_{c}^{2}R^{4}$ for E$\cdotp$J. Here, we have assumed a quasistationary time-step defined by the experimental process $(B_{y},I_{tr})$ which is summarized in the central plot. Top-Left. The magnetic field lines (projected isolevels of the vector potental over the wire cross-section) and their corresponding profiles of current are shown. The consumption of local magnetization currents and the generated field distortions around the superconductor wire are directly visualized. From time to time, straigth isolines in zones free of current are ploted as a consequence of the number of isolevels which has been required to be plotted. They must be understanding only as a visual effect introduced by the graphical processing. Top-Right: Dynamics of the density of power dissipation along the cross-section of the superconducting wire. A clear non symetric distribution of the heat transfer in a cyclic process is always observed. Bottom-Left: The magnetization loop for the superconducting wire in terms of the renormalized electromagnetic sources $(B_{y}/B_{y}^{max}\cdot0.75,0.75)$ is shown. Bottom-Right: The magnetic moment is shown in terms of the virtual time defined by the dynamics of the electromagnetic sources. A low pass filtering effect in the magnetic response as $J_{tr}\rightarrow J_{c}$ is predicted.}

59 — 1108.0975

\caption[]{% Classification into vortex- and streak-dominated families of the simple invariant solutions, in terms of their maximum streamwise and wall-normal r.m.s. velocities, $u'_{max}$ and $v'_{max}$ \citep{jim05}, normalised with the friction velocity $u_\tau$. Solid symbols are classified as vortex-dominated solutions, and open ones as streak-dominated solutions. % The red and blue loops represent the dynamic ($L_x^+\times Re_\tau\times L_z^+=190\times 34\times 130$) and the gentle ($L_x^+\times Re_\tau\times L_z^+ =154\times 28\times 105$) periodic solutions at $Re=400$ in \cite{KawKida01}. % {\color[named]{PineGreen}\trian}, Nagata steady solutions \citep{jim05} for several values of the spanwise wavelength at $Re=400$ and $L_x/h=2\pi$ (upper branch, $L_z^+\times Re_\tau=76-132\times 35-35$; lower branch, $L_z^+\times Re_\tau=53-92\times 24-25$). % {\color[named]{Purple}$\blacksquare$}, autonomous solutions in \cite{JimSim01} ($L_x^+\times L_z^+\times \delta^+=151-189\times 180\times 38-46$, where $\delta$ is the filter height). % {\color[named]{YellowOrange}{\Large \circle}}, \cite{Wally03} travelling waves (upper branch, $Q/\nu=1303$, where $Q$ is the volume flux per unit span, $L_x^+\times Re_\tau\times L_z^+= 387\times 123\times 149$; lower branch, $Q/\nu=1390$, $L_x^+\times Re_\tau\times L_z^+ = 379\times 121\times 146$). % $\square$, \cite{itano01} asymmetric wave for $Q/\nu=4000$, $L_x/h\times L_z/h=\pi\times 0.4\pi$. % The dotted loop is the periodic-like solution of \cite{toh03}. % {\color[named]{SkyBlue}$\blacktriangledown$}, temporal averages of \cite{viswanath07} periodic solutions (cases $P_2-P_6$ in his table~1) for $Re=400$ and $L_x/h\times L_z/h=1.75\pi\times 1.2\pi$. % {\color{green}$\blacklozenge$}, upper branch of \cite{uhlmann10} eight-vortex travelling wave ($Re=1371$, $L_x/h=2\pi$). % {\color[named]{Yellow}\solidcircle}, \cite{wedin04,kerswell05,kerswell07} fourfold rotationally symmetric travelling waves for several values of the streamwise period at $Re=2000$ (${\cal R}_4$ travelling waves in figure 8 of \cite{kerswell05}). % {\color[named]{Lavender}$*$}, twofold rotationally symmetric travelling waves \citep{wedin04,kerswell05,kerswell07} for several values of the streamwise period at $Re=2000$ (${\cal R}_2$ travelling waves in figure 8 of \cite{kerswell05}). % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% We have also shown the turbulent state computed by \cite{kmm} ($L_x^+\times Re_\tau\times L_z^+=2300\times 180\times 1150$) by the symbol {\Large $+$} for comparison. }

\caption[]{ Comparison of the normalised mean velocity profiles between the \cite{KawKida01} time-periodic and turbulent flows. The mean streamwise velocity $\overline{u}^+=\overline{u}/u_\tau$ is shown as a function of the distance from the wall $y^+=yu_\tau/\nu$. {\color{red}{\Large \circle}}, the vortex-dominated periodic solution at $Re_\tau=34$ representing regeneration cycle; {\color{blue}{\Large \solidcircle}}, the streak-dominated gentle periodic solution at $Re_\tau=28$. +, the turbulent solution for the same condition; \solid, the turbulent solution for plane Poiseuille flow at $Re_\tau=590$ \citep{moser99}. {\color[named]{PineGreen}\dashed}, the laminar solution $\overline{u}^+=y^+$. }

\caption[]{ One-dimensional longitudinal energy spectrum $E_\parallel$ for the \cite{veen06} periodic solution in Kida--Pelz high-symmetric flow. The lateral and longitudinal axes are normalised by $(\nu^3/\epsilon)^{1/4}$ and $(\epsilon\nu^5)^{1/4}$, respectively. {\color{red}{\Large \circle}}, the periodic state of high-symmetric flow at $R_{\lambda}=67$; +, the turbulent state of high-symmetric flow; \solidcircle, experimental data of homogeneous shear turbulence at $R_{\lambda}=130$ taken from \cite{Champagne70}. \solid, the high-$R_{\lambda}$ asymptotic form derived theoretically by the sparse direct-interaction approximation \citep{Kida97}. }

\caption{ The intensity of the mean secondary flow $E_{vw}^{1/2}$ (where $E_{vw}$ is the energy of the streamwise-averaged flow in the cross-section, defined in an integral sense) for the duct flow travelling waves of \cite{uhlmann10}, normalized by the bulk velocity $U$ and shown as a function of the Reynolds number. The blue solid line ({\color{blue}\solid}) connects solutions at constant streamwise period $L_x/h=2\pi$; the cross ({\color{blue}$\boldsymbol{\times}$}) corresponds to the flow field visualized in figure~\ref{fig-mu-field}. The closed symbols ($\bullet$) are for long-time averaged turbulent flow in domains with length $L_x/h\approx4\pi$ \citep{uhlmann07}. %; }

60 — 1108.1830

\caption{Room-temperature ($T =$ 295 K) resistance measurements of tunnel junctions with Re (black $\fullsquare$ curve) and Al (red \textcolor{red}{$\fullcircle$} curve) top-electrodes. Resistance measured at 100 nA bias current.}

\caption{Low-temperature ($T$ $\sim$ 50 mK) I-V curves for the Josephson junctions with Re (black $\fullsquare$ curve) and Al (red \textcolor{red}{$\fullcircle$} curve) top-electrodes. The superconducting branch is intentionally suppressed by a magnetic field oriented in the plane of the junction. We measure $\Delta_1+\Delta_2$ = $0.75$ meV for the junction with Re top-electrode, and $\Delta_1+\Delta_2 = 0.45$ meV for the junction with Al top-electrode.}

61 — 1108.1991

\caption{Shown is the two-dimensional, second order structure function of the velocity field $S^v_2(\ell_\perp, \ell_\parallel)$. The offset vector $\boldsymbol{\ell}$ is decomposed into components parallel \emph{($x$-axis)} and perpendicular \emph{($y$-axis)} to the local magnetic field, $\frac{1}{2}(\vsp{B}\left(\vsp{r}+\boldsymbol{\ell}) + \vsp{B}(\vsp{r})\right)$.} \hspace{0.04\textwidth} \label{fig:struct2d} \end{figure} MHD turbulence in the presence of a strong background field, known as Alfv\'{e}nic turbulence, consists of an energy cascade mediated by interacting MHD waves. This situation is treatable with the formalism of wave turbulence \cite[see e.g.][]{EvgenevichZakharov:1992p4576}, whereby the resonant interactions between the MHD ``free particles'' (solutions to the linearized equations) are treated perturbatively. For incompressible MHD, the only free particles are the shear and pseudo Alfv\'{e}n waves. Analysis of the resonant nonlinear interaction of these modes lead to the model of \cite{Goldreich:1995p4506} which predicts a Kolmogorov-like energy spectrum $\propto k^{-5/3}$. Their findings also included the so-called scale-dependent anisotropy with respect to the local mean field, $k_\parallel \propto k_\perp^{2/3}$. This phenomenon is understood geometrically as the exaggerated distortion of eddies, becoming more apparent at smaller scales. The numerical verification was first obtained by \cite{Cho:2000p4507}, who presented two methods of numerical measurement corresponding to the correct geometrical interpretation. Other numerical studies have observed the same scaling in supersonic MHD turbulence \citep{Cho:2003p3040, Beresnyak:2005p4444} and force-free relativistic Alfv\'{e}nic MHD turbulence \citep{Cho:2005p4453}, the latter having received prior analytical treatment by \cite{Thompson:1998p4548}. Here we provide the corresponding measurement for super-Alfv\'{e}nic, compressible relativistic MHD turbulence. We use the same method proposed by \cite{Cho:2000p4507} to measure the scale-dependence of eddy distortions, which relies on the eccentricities of level-surfaces for the second order structure functions \begin{eqnarray} S^v_2(\ell_\perp, \ell_\parallel) &=& \langle |\vsp{v}(\vsp{r}+\boldsymbol{\ell}) - \vsp{v}(\vsp{r})|^2 \rangle \\ S^B_2(\ell_\perp, \ell_\parallel) &=& \langle |\vsp{B}(\vsp{r}+\boldsymbol{\ell}) - \vsp{B}(\vsp{r})|^2 \rangle \end{eqnarray} where $\boldsymbol{\ell}$ is decomposed into cylindrical coordinates oriented along the local mean field, which is defined for each pair of points as $\frac{1}{2}(\vsp{B}\left(\vsp{r}+\boldsymbol{\ell}) + \vsp{B}(\vsp{r})\right)$. In examining the shape of eddies, \cite{Cho:2000p4507} used the structure functions for the velocity and magnetic field $S^v_2$ and $S^B_2$, finding slopes which were on average slightly larger than $2/3$ for $S^v_2$ and slightly smaller for $S^B_2$. As depicted in Figure \ref{fig:magP}, coherent magnetic field structures exist out to only about $1/10$ of the domain size, meaning that $S^B_2$ provides a rather narrow window over which to measure the scaling. However, the velocity field is coherent out the driving scale which affords a broad range of scales over which $S^v_2$ scales robustly. \begin{figure} \includegraphics[width=\ScaleIfNeeded]{\figext{scale-aniso}} \caption{The semi-major ($\ell_\parallel$) and semi-minor ($\ell_\perp$) axes of the elliptical contours obtained from Figure \ref{fig:struct2d} (\emph{circles}), and the best fit line (\emph{solid}) $\ell_\parallel \propto \ell_\perp^{0.84}$. For comparison we provide the prediction of \cite{Goldreich:1995p4506} $\ell_\parallel \propto \ell_\perp^{2/3}$ (\emph{dashed}) and a slope of unity (\emph{dash-dotted}). The vertical line marks the scale $\ell = L/10$ at which the magnetic and kinetic energy are in equipartition.} \hspace{0.04\textwidth} \label{fig:scale-aniso} \end{figure} Indeed, we observe excellent scaling in the structure functions of velocity, indicating that the slope $\ell_\parallel \propto \ell_\perp^{0.84}$ is very robust. In fact, the validity of the fit is valid \emph{at least} between $12\Delta$ and $300\Delta$, the limiting factor in extending the scaling being due to data collection techniques. The fact that the slope is steeper than the $2/3$ means that the eddy distortions depend upon the scale more weakly than in the \cite{Goldreich:1995p4506} model. This may suggest corrections to the cascade dynamics to account for relativistic effects, but we forgo any conclusions on this matter until a detailed comparative study with the equivalent non-relativistic case has been completed. It is also very interesting that the same power law slope of $0.84$ holds above and also below the equipartition scale. Under non-relativistic super-Alfv\'{e}nic conditions, \cite{Cho:2000p4549, Cho:2003p3040} both find similar contour shapes to those shown in Figure \ref{fig:struct2d}. However, in those studies the contour intercepts were not provided, meaning that the precise scaling behavior above and below the equipartition scale is uncertain to us. \cite{Beresnyak:2005p4444} does provide these scalings for trans-Alfv\'{e}nic, supersonic conditions, but only below the equipartition scale. We believe that a rigorous study of scaling above and below the equipartition scale in super-Alfv\'{e}nic turbulence is required in order to isolate the effects of compressibility, substantial magnetic field curvature, and also relativistic effects. \section{Conclusions} \label{sec:conclusions} We have measured spectral and scaling properties of relativistically warm magnetohydrodynamic turbulence in the mildly compressible and super-Alfv\'{e}nic regime. The numerical models were simulated at very high resolution ($1024^3$) using Mara, a new second order Godunov code tuned for accurate and robust evolution of the RMHD equations in three dimensions. Our main production model was driven stochastically at large scales during quasi-steady evolution for about $6$ light-crossing times of the domain. We find that: \begin{enumerate} \item The magnetic energy is amplified from seed fields to $1.5\%$ of the total fluid energy. The scale at which equipartition between magnetic and kinetic energy occurs is between $1/5$ and $1/3$ of the driving scale. \item At $1024^3$ the power spectrum of velocity is dominated by a bottleneck, but not inconsistent with the Kolmogorov prediction of $k^{-5/3}$. The power spectrum of density-weighted velocity $\rho^{1/3} v$ scales $\propto k^{-5/3}$ over moderate wavenumbers, consistent with the simple cascade model of \cite{Kritsuk:2007p3858}. \item About $5\%$ of kinetic energy is in compressive modes. These modes follow a power law over large to moderate scales with index $-1.84$. \item The transverse and longitudinal one dimensional structure functions of velocity are well fit by a power law over moderate to small scales. As a function of the order $p$, the slope of longitudinal velocity fluctuation is well described by the prediction of \cite{She:1994p4537}. Statistically significant deviation is observed for the transverse fluctuation. \item Mild elongation of coherent velocity structures along the local magnetic field is observed. The degree of elongation is scale-dependent, but more weakly than is predicted by \cite{Goldreich:1995p4506}. The scale dependence obeys a power law to high precision above and below the equipartition scale. \end{enumerate} These results suggest that for trans-sonic, super-Alfv\'{e}nic relativistic astrophysical conditions the turbulent cascade dynamics share many similarities with their non-relativistic counterparts. However, the high order scaling relations developed for non-relativistic media, as well as the Alfv\'{e}nic cascade model of \cite{Goldreich:1995p4506} may require modification in order to be applicable to the relativistic astrophysical environments. A detailed comparison between non-relativistic and relativistic MHD models is currently in progress and will form the basis for a future publication. \acknowledgments This research was supported in part by the NSF through grant AST-1009863 and by NASA through grant NNX10AF62G issued through the Astrophysics Theory Program. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center. We would like to thank the Institute for Theory and Computation at the Harvard-Smithsonian Center for Astrophysics for hospitality, where a portion this work was completed. We thank Andrey Beresnyak for helpful suggestions regarding measurement techniques, and Paul Duffell for many useful discussions. We would also like to acknowledge the developers of the Python $\mathtt{numpy}$, $\mathtt{matplotlib}$, and $\mathtt{h5py}$ modules, which were used extensively in our figures and analysis. \bibliographystyle{apj} \bibliography{papers} \end{document} }

62 — 1108.2049

\caption{\label{fig:diag-vs-chol-timing} (Color online) Log-log plots of wall clock times (seconds) vs. basis size $M$, comparing DD (\textcolor{blue}{$\pmb{+}$}) and mCD (\textcolor{xmgrace-green4}{$\blacktriangle$}) of the ERI supermatrix $V_{\mu\nu}$. %The horizontal axis is the number of basis functions, $M$. %The order of the $V_{\mu\nu}$ supermatrix is $M(M+1)/2$. The DD is carried out with 8-thread OpenMP (see text). The mCD timing is obtained from a locally modified MPQC\cite{MPQC-2.3.1} program, for fixed $\delta = 10^{-6}$. No multithreading is used in the mCD calculations. The dashed (dash-dot) lines are linear regressions. The DD slope $\sim M^{5.7}$ is consistent with the expected $M^6$ scaling, while mCD scales as $\sim M^{3.1}$. }

63 — 1108.2974

\caption{Isomorphic representations for SDS and $(\kup,\kdown)=(1,4)$ that produce limit cycles: (a) 3-cycles in $T(8,p)$, (b) 3-cycles in $T(9,p)$, and (c) 4-cycles in $T(9,p)$. \textcolor{red}{to get the multiplicities, we would have to run each of the particular graphs individually.}}

\caption{Isomorphic representations from 1000 random samples of $T(10,p)$ that produced limit cycles greater than length-2 for SDS and $(\kup,\kdown)=(1,4)$ : (a) 3-cycles, (b) 4-cycles (produced by 5 of 6 instances), (c) 4-cycles (produced by 1 of 6 instances), and (d) 5-cycles. \textcolor{red}{to get the multiplicities, we would have to run each of the particular graphs individually.}}

64 — 1108.4186

\caption{(Color online) Low-temperature polarization as a function of magnetic field for the different $P$ and $H$ orientations considered. The inset shows the variation of the temperature onset of polarization $T_P^*$ for ($P||c$, $H||a$) (\textcolor{red}{$\bigtriangledown$}) and ($P||a$, $H||c$) (\textcolor{green}{$\bigtriangleup$}), as well as the temperature $T_M^*$ (\textcolor{black}{$\bullet$}) below which a drop/step is observed in the magnetization recorded with $H||c$.}

65 — 1108.5567

\caption{Visualization of parse tree~\eqref{ex:dbj1} for sentence~\eqref{eq:s1} using \idpdraw.}

66 — 1108.5815

\caption{Timings on CPU for Laplace kernel (potential+force), \textcolor{update}{treecode with $p=5$ and $p=8$ for FMM and hybrid}---(A) treecode, FMM and hybrid method, with particles randomly placed in a cube; (B) FMM and hybrid method with different values of $N_{crit}$; (C) particles randomly placed on a spherical shell.}

\caption{Breakdown of the run time on GPU for the three methods. $N=10^7$ particles randomly placed in a cube, Laplace kernel potential+force, \textcolor{update}{$p=5$ for the treecode and $p=8$ for FMM and hybrid.} }

67 — 1109.0188

\caption{(a) Autocorrelation function of the vertical component of the micro-bubble acceleration for the different $\mathrm{Re}_{\lambda}$ measured. The correlation of the micro-bubble acceleration persists longer with increasing Reynolds number. (b) The decorrelation time $T_D/\tau_{\eta}$ of the autocorrelation function for the three components of the micro-bubble acceleration as a function of $\mathrm{Re}_{\lambda}$. The decorrelation time increases with the turbulent intensity. %The micro-bubble acceleration is nearly isotropic as the three components have very similar decorrelation times. A linear fit considering the experimental data of $\mathrm{a}_z$ gives: $T_D/\tau_{\eta}=0.00046\mathrm{Re}_{\lambda}-0.054$. In the inset, we show also the result of Volk \etal~\cite{Volk2008a} at a very high $\mathrm{Re}_{\lambda}$=850 ({\color{-green}$\blacklozenge$}), their experimental point agrees with the trend of increasing decorrelation time with turbulent intensity. The linear fit obtained with our experimental data extrapolates a value of $T_D/\tau_{\eta}=0.27$ at $\mathrm{Re}_{\lambda}$=850, which is slightly higher than their experimental value of $T_D/\tau_{\eta}=0.25$.}

68 — 1109.0440

\caption[]{\textbf{Results.} \textbf{a},~Visibility as a function of pump power. The visibility is approximately constant with an average of $96.5\pm 1.2\%$ (green shaded region). The inset shows the visibility curves at 16~mW measured with detectors~1 (\bluecircle) and~2 (\redsquare) (15 minutes acquisition time per point; the error bars are smaller than the symbols). The different amplitudes result from non-uniform loss after recombination of modes $A$ and $B$ on the beamsplitter. The visibilities agree within uncertainties in the fits. \textbf{b},~Zero-time cross-correlation $\bar{g}_{s,i}$ as a function of pump power. The decreasing values agree well with a theoretical model (shaded region; see \apdxname). \textbf{c},~Lower bound on the concurrence estimated using the cross-correlation measurement the as a function of pump power (\bluecircle), calculated using Eq.~\ref{C} and~\ref{p11}, and the mean visibility of \textbf{a}. The concurrence decreases with pump power, as expected, but remains positive up to 16~mW. The shaded region corresponds to the model in \textbf{b}. All measured values of $p_{10}$, $p_{01}$ and $p_{11}$ are given in the \apdxname. All values are based on raw counts. Uncertainties are obtained assuming Poissonnian detection statistics. The lower bound on the concurrence at 16~mW obtained from measured threefold coincidences using either the maximum likelihood estimation (\greendiamondempty) or the conservative estimation (\greendiamond) are also shown (they are horizontally offsetted for clarity).}

69 — 1109.1350

\caption{\label{fig:TypeAB} (Color online) Twofold structurally degenerate inequivalent Type IIA and Type IIB FCC structures. The four colors - \textcolor{black}{\bf black}, \textcolor{red}{{\bf red}}, \textcolor{blue}{\bf blue}, and \textcolor{green}{\bf green} represent the four sublattices respectively. The $+$ and $-$ denote the up and down collinear spin configurations respectively. Spin wave theory upto quartic order in the ratio of interaction strengths show that quantum fluctuations select the Type IIA structure as having the lower energy ~\cite{HarrisFCC}. We focus on this spin arrangement in this paper.}

\caption{\label{fig:fcc} (Color online) Four sublattice formulation of the AF FCC lattice. The sublattices are indicated by - \textcolor{black}{\bf black} (1), \textcolor{red}{{\bf red}} (2), \textcolor{blue}{\bf blue} (3), and \textcolor{green}{\bf green} (4) colors. The corresponding sublattice number is indicated in the parenthesis. The intra-sublattice nearest neighbor vector is given by ${\bf \Delta}$ and inter-sublattice vectors are given by $\Vec{\delta}_{12},\Vec{ \delta}_{13}$ and $\Vec{\delta}_{14}$. The choice of coordinate axis is as shown in the figure.}

70 — 1109.2301

\caption{The $\mu$-diagram vs. the number of particles $N$ for $u_C = 0.5$. The \textcolor{blue}{blue} points correspond to single-particle states, the \textcolor{red}{red} points to two-particle states, the \textcolor{brown!60!black}{brown} points to three-particle states, and the only \textcolor{black}{black} point to four particles. The bias window is $\Delta\mu=\mu_L-\mu_R$. The small bias window is $\Delta\mu=3.20-2.98=0.22$ (\textcolor{green!30!black!75}{green}). The large one is $\Delta\mu=3.30-2.90=0.40$ (\textcolor{violet!75}{violet}). Both bias windows include the ground state of three particles, but also excited one and two-particle states. So the expected number of electrons in the steady state is slightly below three. }

\caption{The time dependent charge for $\mu_L = 3.20$, $\mu_R = 2.98$. The contact functions $\chi_{L,R}$, given by Eq.\\ref{eq:turnstile}, are shown at the bottom. (a) Left lead at site 1 (\textcolor{blue}{blue}), right lead at site 10 (\textcolor{red}{red}). (b) Left lead at site 1, right lead at site 3. The total charge is shown by the continuous \textcolor{brown}{brown} line. The dashed lines show the population of the states with \textcolor{black}{1}, \textcolor{violet!60!black}{2}, and \textcolor{olive}{3} electrons (indicated as $Q_1$, $Q_2$ and $Q_3$ on the top of the figures). }

\caption{A snapshot of the charge distribution along the 10 sites, for $\mu_L = 3.20$, $\mu_R = 2.98$. The time is $t=400$. The position of the right contact is shown on the horizontal axis in \textcolor{red}{red}. (a) Left 1 Right 10. (b) Left 1 Right 3.}

71 — 1109.2317

\caption{\revised{Comparison among traditional erasure codes, regenerating codes and self-repairing codes (derived theoretically in \cite{OD-infocom}):} Average traffic normalized with $B/k$ per lost block for various choices of $x$ ($B$ is the size of the stored object) for \colortext{(n=31,k=8)} encoding schemes. For parallel repairs using erasure codes the traffic is $k=8$ (not shown). \colortext{The SRC code parameters are denoted as SRC(n,k).}}

72 — 1109.3066

\caption{\label{S0_BT}Temperature and magnetic field dependence of the spin-Seebeck coefficient (in units of $k_{B}/\left|e\right|$) of a noninteracting quantum dot. The maxima at $\tilde{B}=\Gamma$, $T=0.34\Gamma$ and $T=0.34\tilde{B}$ are indicated with dashed lines. (a) Exact result, Eq.~(\ref{eq:exact_NI}). (b) Interpolation formula, Eq.~(\ref{eq:interp_NI}).} \end{figure} In the absence of interaction the impurity Green's function is $G_{\sigma}\left(\omega\right)=1/\bigl(\omega-\sigma\tilde{B}+i\Gamma\bigr)$. We introduced the energy level shift $\tilde{B}$ due to the magnetic field. In the non-interacting case, $\tilde{B}=\frac{B}{2}$. From Eqs.~(\ref{eq:S}) and (\ref{eq:exact_NI0}) \begin{equation} S_{s}={\cal S}\bigl(\tilde{B}\bigr)-{\cal S}\bigl(-\tilde{B}\bigr)=2{\cal S}\bigl(\tilde{B}\bigr).\label{eq:exact_NI} \end{equation} For $T\ll\max\left(\Gamma,\tilde{B}\right)$ the spin-Seebeck coefficient is proportional to the temperature, \begin{equation} S_{s}=\frac{4\pi^{2}}{3}\frac{\tilde{B}T}{\tilde{B}^{2}+\Gamma^{2}}, \end{equation} a results which can also be derived by performing the Sommerfeld expansion of the transport integrals in Eq.~(\ref{eq:transint}). It increases linearly with the field in low magnetic fields, $\tilde{B}\ll\Gamma$, \begin{equation} S_{s}=\frac{4\pi^{2}}{3}\frac{\tilde{B}T}{\Gamma^{2}},\label{eq:nonint_lowBT} \end{equation} and is inversely proportional to the field in high magnetic fields, $\tilde{B}\gg\Gamma$, \begin{equation} S_{s}=\frac{4\pi^{2}}{3}\frac{T}{\tilde{B}}.\label{eq:nonint_lowT} \end{equation} It reaches its maximal value at $\tilde{B}=\Gamma$. For $\tilde{B}\ll\max\left(\Gamma,T\right)$ the spin-Seebeck coefficient is proportional to the magnetic field: \begin{equation} S_{s}=\frac{2\tilde{B}}{T}\left(1+\frac{\Gamma}{2\pi T}\frac{\psi^{\prime\prime}\left(\frac{1}{2}+\frac{\Gamma}{2\pi T}\right)}{\psi^{\prime}\left(\frac{1}{2}+\frac{\Gamma}{2\pi T}\right)}\right).\label{eq:nonint_lowBx} \end{equation} In the low temperature limit, $T\ll\Gamma$, where we recover Eq.~(\ref{eq:nonint_lowBT}), it increases linearly with the temperature, reaching its maximal value at $T=0.34\Gamma$. It is inversely proportional to the temperature in the $T\gg\Gamma$ limit: \begin{equation} S_{s}=\frac{2\tilde{B}}{T}.\label{eq:nonint_lowB} \end{equation} As shown in Fig.~\ref{S0_BT}, an approximation that smoothly interpolates between the regimes of Eqs.~(\ref{eq:nonint_lowBT}), (\ref{eq:nonint_lowT}) and (\ref{eq:nonint_lowB}), \begin{equation} S_{s}^{-1}=\left(\frac{4\pi^{2}}{3}\frac{\tilde{B}T}{\Gamma^{2}}\right)^{-1}+\left(\frac{4\pi^{2}}{3}\frac{T}{\tilde{B}}\right)^{-1}+\left(\frac{2\tilde{B}}{T}\right)^{-1},\label{eq:interp_NI} \end{equation} is qualitatively correct also at $\tilde{B},T\sim\Gamma$. According to this interpolation formula, the spin-Seebeck coefficient at $\tilde{B}=\Gamma$, $T=0.34\Gamma$ where the maxima merge is $S_{s}=1.45$, which compares well with the exact result of Eq.~(\ref{eq:exact_NI}), $S_{s}=1.61$. From this point, the high spin-Seebeck coefficient ridge of $S_{s}\sim1$ continues to higher temperatures and fields along the $T=0.34\tilde{B}$ line. Note also that at the ridge the magnetic field and the temperature are of the same order of magnitude. \subsection*{Away from the particle-hole symmetric point, $\boldsymbol{\delta\ne0}$} \begin{figure} \begin{centering} \includegraphics[width=1\columnwidth]{fig3} \par\end{centering} \caption{\label{S0_BT_away}Temperature and magnetic field dependence of the spin-Seebeck coefficient (in units of $k_{B}/\left|e\right|$) of a noninteracting quantum dot away from the particle-hole symmetric point for $ $(a) $\delta=\Gamma/2$, (c) $\delta=\Gamma$ and (e) $\delta=2\Gamma$. The corresponding electrical voltages, $-eV/\Delta T$, required for electrical current to vanish are shown in (b), (d) and (f). Along the dashed lines in (a), (c) and (e) the required electrical voltage is zero.}

73 — 1109.4423

\caption[]{{\it HST}/WFC3 color-magnitude diagrams for the stellar field ($512 \times 512\ \mathrm{pixels}$, i.e. $20\arcsec \times 20 \arcsec$) around the ULX. Padova isochrones for stars of different ages are overplotted. Typical photometric errors are also plotted. Data have been corrected for an extinction of E$(B-V) = 0.08$ mag, the bar at the bottom right corner illustrating this effect. Upper left panel: Color-magnitude diagram in the (F225W,F336W) system. Upper right panel: Color-magnitude diagram in the (F336W,F547M) system. Lower left panel: Color-magnitude diagram in the (F547M,F845M) system. The same isochrones are plotted in the three panels, i.e for 5, 10, 20, and 50 Myr at Z = 0.008. Filled green circles are stars in the field, red diamonds are stars located in a rectangle of $5.5 \times 4.0 \arcsec$ coincident with a nearby OB association. The red star is the position of the ULX counterpart. In the bottom panel, we show as orange X's stars from the OB association that are in common with the upper panels. \red{The smaller number of stars in common with the upper panels is simply due to the fact that we mainly see redder stars in the F845M filter} that are not strong UV emitters (and vice-versa in the top panels where the strong UV emitters are not seen in the redder filters). }

\caption[]{Left: best fits for the simultaneous X-ray - UV/optical/NIR data of NGC~5408~X-1. The three plots correspond to the three independent observations. In each plot, the two accretion disk models (see text) extrapolated from the X-ray data fit to lower energies are shown in green (dot-dashed line) and in blue (dashed line)\red{, respectively corresponding to the disk model associated with a Comptonized component and the disk model associated with a power law component}. They underestimate the UV/optical/NIR fluxes. The best fitting model is an irradiated disk (red line) that explains both the high energy and low energy data. This supports the interpretation that a standard, thin disk is present in NGC~5408~X-1. The spectrum of a B0I supergiant star (from the Castelli and Kurucz stellar atmosphere models, Castelli et al. 2004) is overplotted in orange. It fits the {\it HST} data as well as the irradiated disk model. Note that the apparent gap between $\sim 100$ and $\sim 1000\ \mathrm{\AA}$ is due to the extinction. Right: required flux needed to explain the {\it HST} measurements, after subtraction of \red{the two extrapolated disk components seen in the left panels (the same color/linestyle is used to differentiate the two disk models)}. }

74 — 1109.5825

\caption{(Color) Forcing parameter space. The points represent experiments where: $\blacksquare$ no lock-in is observed; {\color{blue}$\bigstar$} lock-in is observed. The blue line is the lock-in threshold estimated from the present experiments. The red line stands for the observations of \cite{thiria2007}. Striped region indicates a maximum drag reported by \cite{bergmann2005}.}

75 — 1110.0034

\caption{Compton composite ``classical" depictions showing {\color{red}{protons}} and {\color{green}{electrons}}.}

76 — 1110.1034

\caption{\textcolor{red}{Simulated photometric and RV signatures of a single spot (top and bottom respectively). The solid black line shows the output of our simple model for a fairly large, dark, equatorial spot ($c=0$, $\alpha = 10\,^\circ$, $\delta=0$) on a star with $P_{\rm rot}=5$\,days,$i=90\,^\circ$, $\delta V_{\rm c}=200$\,m\,s$^{-1}$ and $\kappa=10$ (see text for details). The solid cyan line is the same, but for a higher latitude spot on an inclined star ($\delta=60\,^\circ$, $i=70\,^\circ$). For comparison, the dashed lines show the same spots modeled with the formalism of \protect\citet{dor87} (stellar linear limb-darkening parameter $u_\star=0.5$). For simplicity, we have omitted the Dorren formalism for the high-latitude case in the bottom panel. Instead, the black dash-dot line shows the equatorial spot simulated with our simple model, but without the convective blue-shift effect ($\delta V_{\rm c}=0$). }}

\caption{Lomb-Scargle periodograms of the observed RV time-series for HD\,189733 after subtracting only the planetary orbit (solid black line) and after removing the simulated activity signal also (\textcolor{red}{dashed} grey line). Both RV datasets have the same time sampling, including only SOPHIE data points simultaneous with the MOST observations, and excluding the 4 data points between the vertical dashed lines in Figure~\protect\ref{fig:hd189t}. The rotational frequency and its \textcolor{red}{harmonics} are indicated by the vertical \textcolor{red}{dotted} lines. Following \protect\citet{boi+09}, we took the rotation period to be 11.953\,days.}

\caption{Example Kepler Quarter 1 light curve before and after applying our systematics correction. The original, raw time-series is shown in grey, the corrected time-series in black, and the corrected time-series without the transits (used to estimate the RV modulations) in red. \textcolor{red}{Note that the red line completely overlaps with the black, except during the transits.} This example is KID 3642741 (KOI 242).}

77 — 1110.2037

\caption{(Color online). Plot of running of the quark mass as a function of energy scale for different $r$. The blue ({\color{blue}---}) is the running of quark mass from QCD. The green ({\color{green}.......}) is for $r=-0.29\,(b=0.52)$, the orange ({\color{BurntOrange}--- --- ---}) is for $r=-0.53\, (b=0.60)$, and the red ({\color{red}--- -- --- -- ---}) is for $r=-0.75\, (b=\sqrt{2/3})$.}

\caption{(Color online). Plot of running of the quark condensate as a function of energy scale for different $r$. The blue ({\color{blue}---}) is the running of quark mass from QCD. The green ({\color{green}.......}) is for $r=-0.29 \,(b=0.52)$, the orange ({\color{BurntOrange}--- --- ---}) is for $r=-0.53\, (b=0.60)$ and , the red ({\color{red}--- -- --- -- ---}) is for $r=-0.75\, (b=\sqrt{2/3})$. }

78 — 1110.2495

\caption{(Color online) (a) The variational phase diagram in a magnetic field. AFM2 denotes a two-sublattice ordered N\'eel phase, while FM stands for ferromagnet. Note the presence of a magnetization plateau at 1/2 which is separated from AFM2 by a supersolid phase. (b) Schematic plot of the anticipated phase diagram based on exact diagonalization calculations. AFM3 (AFQ3) is a three-sublattice ordered antiferromagnetic (antiferroquadrupolar) phase. The upper boundary of the 1/2 plateau is a result of finite-size scaling. The dashed line is estimated within flavor-wave theory (see text).\label{fig:PDh}}

\caption{(Color online) Spin and quadrupole structure factor from exact diagonalizations of square clusters with periodic boundary conditions. The geometry of the clusters can be found in Fig.~{\color{red} 72} of Ref.~\onlinecite{Dagotto1994}. Open (closed) symbols stand for the spin (quadrupole) structure factor. \label{fig:SqQq}}

79 — 1110.2700

\caption{Two examples of the variability of the pulse shape, left an example early in the observation, right one close to the end of the observation. Shown in black is the 1\,sec light curve, in red a 20\,sec light curve. The blue profile is the average pulse profile over the complete observation. Note that the$y$-axes are not the same scale.}

\caption{Small part of the light curve between 0.3--12.0\,keV, showing a vanishing of the pulsations. In black the light curve with 5\,sec resolution is shown, in red with 40\,sec resolution. Superimposed in blue is the average pulse profile for the whole observation.}

\caption{\textit{a)}: Phase and time averaged spectrum in the 2--10\,keV energy range of EPIC-pn. The components of the spectral model are shown in different colors, see text for a detailed description.\textit{b)} Residuals with neither an absorption feature nor an Cr K$\alpha$ line around 5.4\,keV.\textit{c)} Residuals with a Cr K$\alpha$ line, but without an absorption feature. \textit{d)} Residuals with an absorption feature, but without a Cr K$\alpha$ line. \textit{e)} Residuals without a Ni~K$\beta$ line. \textit{f)} Best fit residuals. }

\caption{Time series of the spectral parameters of the spectra of individual pulses. \textit{a)} Light curve between 0.5--10\,keV,\textit{b)} continuum norm in photons\,keV$^{-1}$\,s$^{-1}$\,cm$^{-2}$, \textit{c)} main absorption column in $10^{22}$\,\atmcmsq, \textit{d)} covering fraction, \textit{e)} gainshift slope, and \textit{f)} \redchi of the best fit.}

80 — 1110.2711

\caption{\label{fig_rw_javax}Community structure of Java software network revealed with \mpa~(b). Only communities with more than $24$ nodes are shown, still, the structure contains $1020$ nodes and $4184$ links. Link colors correspond to high-level software packages---{\color{brown}\texttt{javax.swing}}, {\color{blue}\texttt{javax.management}}, {\color{yellow}\texttt{javax.xml}}, {\color{green}\texttt{javax.print}}, {\color{lightgray}\texttt{javax.naming}}, {\color{darkgray}\texttt{javax.lang}} and {\color{black}other}---while each dot was enlarged five times for better visibility.}

81 — 1110.2815

\caption{\label{fig:residual-DARM}Calibrated DARM displacement (upper traces) and residual motion (lower traces), shown as amplitude spectral density (\full, in meters per $\sqrt{\textrm{Hz}}$) integrated root-mean-square displacement (\dashed, in meters). The RMS DARM displacement above 0.07 Hz is reduced by approximately eight orders of magnitude by the control system in order to remain sufficiently near the operating point. The residual motion sets a lower limit for the DARM offset in DC readout to avoid fringe-wrapping.}

\caption{\label{fig:shot-noise-comparison} Achieved sensitivity (noise floor) amplitude spectral density for the Livingston (left) and Hanford (right) detectors (\full), with expected shot noise limit (\dashed). The Livingston spectrum is from 2010-08-11 06:34 UTC and the Hanford spectrum is from 2010-07-04 08:52 UTC. The estimated uncertainty in the modeled curve is approximately~ $\pm15\%$. }

82 — 1110.3624

\caption{(Color online) Field-cooled magnetization data recorded on cooling (FCC) as a function of temperature for a natural mineral (\textcolor{black}{$\blacklozenge$}) and synthetic (\textcolor{blue}{$\bullet$}) sample melanostibite measured in a) 0.02 mT and b) 1 T. The inset shows the specific heat of the synthetic sample measured in zero magnetic field.}

\caption{(Color online) a) Magnetic field dependence of the magnetic entropy change $\Delta S_M (H)$ for the natural mineral at different temperatures. b) $\Delta S_M (H)$ for the natural mineral (\textcolor{black}{$\blacklozenge$}) and synthetic (\textcolor{blue}{$\bullet$}) melanostibite at their respective transition temperature and the $H^{2/3}$ power law dependence (dashed lines).}

\caption{(Color online) The magnetic entropy change $\Delta S_M$ under 1 T field change as derived from specific heat (\textcolor{red}{$\circ$}) and $M$ vs. $H$ (\textcolor{black}{$\star$}). $-d$$M$/$d$$T$ $\times$ $H$ was derived from $M$ vs. $T$ in 1 T applied field(\textcolor{blue}{$\diamond$}).}

83 — 1110.3646

\caption{\textcolor{red}{The $M$-legged, $\mathcal{N}$-runged quantum spin-1/2 ladder in the form of a \textit{bipartite lattice}}. The solid circles are of sublattice $A$, while the hollow ones are of $B$. The red arrows (solid) show the nearest-neighbor dimer states ($|(a_i,b_j)\rangle$) from a site in sublattice $A$ to another in $B$. The arrows (dashed) at the boundary indicate that the lattice is with periodic boundary condition. The dashed lines indicate that the number of legs and rungs can be extended beyond the illustrated number of sites. %show the considered periodicity of the lattice. The figure also shows the two-site ladder state, $|2\rangle$, and the one-site rung state, $|1\rangle$, which will be profusely used in the iteration method presented.}

84 — 1110.3734

\caption[For LoF]{Radial temperature profiles in the cool solvent around a hot nanoparticle of radius $R\approx 5$ for the particle temperatures $T_\mathrm{p}=1.00$~{\Large\color{myblue}$\bullet$}, $1.25$~{\color{mypurple}{\small $\blacksquare$}}, $1.50$~{\color{myokker}{\Large$\diamond$}}, $1.75$~{\color{mygreen}$\blacktriangle$}, $2.00$~{\color{mymarine}$\blacktriangledown$}, $2.25$~{\color{mypurple}{\Large $\circ$}}. Dashed and solid lines are fits obtained from Fourier's law for a thermal conductivity $\lambda\,$=constant (dashed), as assumed in theory, and $\lambda(T)\propto 1/T$, as inferred from the fit to independent simulations of an isothermal bulk fluid (inset), respectively.}

\caption{Sample points of the radial temperature and density profiles $T(r)$ and $\rho(r)$ around a hot nanoparticle maintained at various temperatures $T_\mathrm{p}$, in the $T-\rho$ plane. The filled circles ({\Large\color{mycyan}$\bullet$}) correspond to states for which the viscosity $\eta(T)$ of the bulk Lennard-Jones fluid was determined in separate simulations. The inset compares the latter($\square$) to the empirical formula from \fref{eq:etaT_fit} ({\bf ---}) and alternative expressions proposed in the recent literature \cite{Rowley:1997} ({\color{mypurple}$\mathbf{--}$}) \cite{Galli'ero:2005} ({\color{mygreen}$\mathbf{-\cdot-}$}).}

\caption{Effective temperatures of translational hot Brownian motion. \emph{Simulation:} $\THBM$ ({\Large \color{myblue}$\bullet$}) from the generalised Einstein relation, \fref{eq:GER}; apparent equipartition temperature $T_\mathrm{k}$ ({\color{mypurple}{\Large $\circ$}}) for the particle velocity; solvent temperature at the particle surface (dotted) \emph{Theory:} $\THBM$ according to \fref{eq:T_HBM_Manuel} (solid line) and the previous thermodynamic estimate, eq.~(22) of Ref.~\cite{Rings:2011} (dashed line), both evaluated (assuming an incompressible fluid) with the exact numerical differential shell method \cite{Rings:2011}.}

85 — 1110.4890

\caption{{\bf Dependencies of the half-maximum shift on laser power and detuning.} $\Delta_{HM}$ versus $v_{00}^s$ laser power for $\Delta v_{00}^s$= -340 MHz \small(\footnotesize${\color{Orange}\blacksquare}$\small), -260 MHz \small(\large{\color{green}$\bullet$}\small), -180 MHz ({\color{blue}$\blacktriangle$}) and associated fits.}

86 — 1110.5494

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

87 — 1110.6043

\caption{\footnotesize Major \gray flares detected by the \grid in the period November 2007 - July 2009. All detections have a significance above $3\sigma$ ($\sqrt{TS} \geqslant 3$). \textit{Column one}: period of detection in MJD; \textit{Column two}: significance of detection; \textit{Column three}: photon flux.}

\caption{\footnotesize Multi-frequency light curve of Cygnus X-3 from 2007-December-12 to 2009-September-26 (MJD: 54450-55100). From top to bottom: \textbf{radio} flux density [RATAN-600 (2.15, 4.8, 11.2 GHz) and AMI-LA (15 GHz)], \textbf{soft X-ray} count rate [\textit{RXTE}/ASM (3-5 keV)], \textbf{hard X-ray} count rate [\textit{Swift}/BAT (15-50 keV)] and $\boldsymbol\gamma$\textbf{-ray} photon fluxes [\grid (above 100 MeV)]. In the bottom panel gray regions represent the AGILE pointing at the Cygnus region; \textit{magenta} points are the \gray flares with $\sqrt{TS}\geqslant3$ (major \gray flares, see Table \ref{cyg_x3_all_flares}), \textit{black} points are the \gray detections with $2 \leqslant \sqrt{TS} < 3$ and \textit{dark-gray} arrows are $2\sigma$ the upper limits related to $\sqrt{TS} < 2$.}

\caption{\footnotesize Schematic representation of the evolution of Cygnus X-3 through its radio and X-ray state. The red stars mark the approximate position of the \gray flares detected by the \grid, occurring in pre-quenching/pre-flaring radio states.}

88 — 1110.6629

\caption{Viscoelasticity $\phi(\omega)$ of cantilevers {\color{blue}$\bf cAu$}, {\color{red}$\bf cAl$} and {\color{black}$\bf cPt$} reconstructed from the noise spectrums measured in vacuum.\label{Fig:metals}}

89 — 1111.0697

\caption{ Historical (2007 - 2008, black symbols and inset; data from \cite{Vercellone2008:3C454_ApJ,Vercellone2008:3c454:ApJ_P1,Donnarumma2009ApJ_P2,Vercellone2010ApJ_P3}) and super-flares (2009 - 2010, blue symbols; data from \cite{Striani2010ApJ_3c454} and \cite{Vercellone2011:ApJL_3C454_nov2010}, respectively) AGILE \gray{} light-curves for $E>100$\,MeV in units of$10^{-6}$\,\phcmsec{}. During the 2007 - 2008 campaign the \gray{} flux level was for most of the time higher than the maximum flux detected by EGRET (dotted gray line). }

\caption{ $R$-band, 230~GHz, and \gray{} light-curves (top, middle and bottom panel, respectively) covering the period 2007 July -- 2009 January. The AGILE light-curve has a time-bin of 1-week. Adapted from \cite{Vercellone2010ApJ_P3}. }

\caption{ \swi/XRT photon index as a function of the 2--10\,keV flux. Open circles and squares refer to photon counting and windowed data, respectively. Black, green and red symbols represent data accumulated during the 2007--2009 monitoring campaign\cite{Vercellone2010ApJ_P3}, the 2009 December super-flare \cite{Pacciani2010ApJ_3c454}, and the 2010 November super-flare \cite{Vercellone2011:ApJL_3C454_nov2010}, respectively. }

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

\caption{From top to bottom: \gray{} ($E>100$\,MeV), X-ray (2--10\,keV), UV ($w1$, $m2$, w2), and optical ($R$) light-curves collected before, during, and after the 2010 November 20 (MJD~55520) \gray{} super-flare. The ``\gray-orphan'' optical-UV flare is visible at MJD~55510.}

90 — 1111.2813

\caption{Charge exchange cross sections for \muplus\in helium, velocity scaled from proton cross sections in~\cite{nak87} resulting in final states with \mucplus\(large dashes), \muczero\(small dashes), and \mucminus\(dots). The two-state charge-change-cycle cross section is shown in solid\ifcolor \red\fi.\label{fig:charge_exchange_xs}}

\caption{Effective charge (line) for \muplus\in He, H, Ne, and C for the three-state (solid\ifcolor\black\fi) and two-state (dashed\ifcolor\red\fi) systems. The data points show the equilibrium charge state fractions taken from \cite{RevModPhys.30.1137}\label{fig:effective_charge}.}

\caption{Stopping power of helium on \muplus\(solid) and accelerating power of a uniform constant electric field of strength $E$ on \muplus\in helium with (large dashes\ifcolor, red\fi) and without (\ifcolor\else horizontal, \fi small dashes\ifcolor, black\fi) accounting for effective charge for two values of $eE$ (as indicated on the right axis).\label{fig:dedx_with_effcharge_He}}

91 — 1111.2920

\caption{Representation of a nanostructured channel filter as modelled in the present Letter. The radiuses \rcero\and\rhocero\correspond to the average dimensions of the bare channel and impurities. The effective radiuses\re\and\rhoe\vary as trapped impurities cover the inner wall, via their dependences on, respectively, the areal density\ene\of trapped impurities and on the areal density\ze\of effective charge of the inner wall. This\ze\reflects that exposed charges in a nanostructured surface attract the impurities in the fluid and also constitute binding anchors for those impurities. It is expected to diminish as impurities cover the surface, for which we assume the simple$\ze(\ene)$ dependence plotted at the right (although it can be easily generalized to more complex functionalities, see main text).}

\caption{~(a):~Results, obtained by integrating \eq{diferencialred}, for the areal density of trapped impurities in units of the saturation value, $\enered=\ene/\enesat$ (continuous line), as a function of time in units of the half-saturation value, $\tred/\tredhalfsat$ with \tredhalfsat\defined by$\enered(\tredhalfsat)=0.5$. The parameter values used are (see main text for details): $\rhoesat=\rhocero=13$\AA, $\rhoeclean=20\rhocero$, $\resat=450\rhocero$, $\reclean=500\rhocero$, $\enesat=5/\rhocero^2$, $\omegabindingsat=0$, $\omegabindingclean=0.5/10^3\rhocero$, and the initial value was $\enered(0)=0$. We also show the results obtained from the clean-regime approximation, \eq{clean-1} (dashed line), from the half-saturation approximation, \eq{sat-1} (dotted line), and for the saturation limit, $\ene=\enesat$ (dot-dashed line). (b):~Filtration capability $\dd\enered/\dd\tred$ (continuous line) and fluid flow \flujo\(dashed line), normalized to their values at $t'=0$ and as a function of $\tred/\tredhalfsat$, obtained again from \eq{diferencialred} and the same parameter values as in Fig.~2a. Note that the horizontal axes of these plots are logarithmic, so that the half-saturation regime is well longer than the clean regime. The hydrodynamic flow through the channel is reduced only about 30\% during its working lifetime.}

92 — 1111.4141

\caption{\Measured\textsuperscript{131}I and \textsuperscript{137}Cs air activity concentrations during the first 62 d of air sampling. Filters collected air particles for approximately 24 h.}

\caption{\Assay results from filters measured at KURF. The filter efficiency for\textsuperscript{132}Te is assumed to be 99.98\%. All activity concentrations are in units of mBq/m\textsuperscript{3}.}

93 — 1111.4460

\caption{As an example, consider a scenario with $n=2$ and $m=3$. The arms $U = \left\{ u_1, u_2, u_3 \right\}$ and the preference vector $z^{*}$ are located at the indicated points. The shaded region is $A$; the boundary of $A$ is formed by the perpindicular bisectors of the segments $u_3, u_1$ and $u_1, u_2$.} \label{fig:three_phase_A} \end{figure} \begin{lem}\label{lemma bad epoch} The probability that epoch $l$ is a bad epoch is upper-bounded by $$2n \cdot \exp\left\{ -2l\cdot f \left( - \left\Vert z^{*}\right\Vert \right) \right\}.$$ \end{lem} Proof: In order for epoch $l$ to be a good epoch, we have the condition that $\hat{\alpha}_{u,l} \in (0, 1) \\forall u \in \Sigma$. Consider the condition for a bad epoch: \begin{align*} & \exists u \in \Sigma \mbox{ s.t. } \hat{\alpha}_{u,l} \notin (0, 1) \\ & \implies \exists u \in \Sigma \mbox{ s.t.} \left| \hat{\alpha}_{u,l} - \frac{1}{2} \right| \ge \frac{1}{2} \\ & \implies \exists u \in \Sigma \mbox{ s.t.} \left| \hat{\alpha}_{u,l} - \alpha_u^* \right| \ge \frac{1}{2} - \left| \alpha_u^* - \frac{1}{2} \right| \\ & \implies \exists u \in \Sigma \mbox{ s.t.} \left| \hat{\alpha}_{u,l} - \alpha_u^* \right| \ge \frac{1}{2} - \max_{v \in \Sigma}\left| \alpha_v^* - \frac{1}{2} \right|. \end{align*} Note that $$\max_{v \in \Sigma}\left| \alpha_v^* - \frac{1}{2} \right| \le f \left( \max_{u \in \Sigma} \left\Vert u \right\Vert \cdot \left\Vert z^{*}\right\Vert \right) - \frac{1}{2} .$$ Then, applying the union bound and Chernoff bound, it follows that \begin{align*} & P \left( \exists i \in \Sigma \mbox{ s.t. } \hat{\alpha}_{i,l} \notin (0, 1) \right) \\ \le & 2n \cdot \exp\left\{ -2l\cdot \left[ 1 - f \left( \max_{u \in \Sigma} \left\Vert u \right\Vert \cdot \left\Vert z^{*}\right\Vert \right) \right] \right\} \\ = & 2n \cdot \exp\left\{ -2l\cdot f \left( - \max_{u \in \Sigma} \left\Vert u \right\Vert \cdot \left\Vert z^{*}\right\Vert \right) \right\}. \end{align*} Let $k_1 = 2 f \left(\max_{u \in \Sigma} \left\Vert u \right\Vert \cdot \left\Vert z^{*}\right\Vert \right)$. Thus, the probability that epoch $l$ is a bad epoch is then upper-bounded by $$2n \cdot \exp ( -k_1 l ) .$$ \begin{flushright} $\square$ \end{flushright} \begin{lem}\label{lemma finite per timestep phase 2} For the Two-Phase Algorithm on finite arms, for a given choice of scheduling function $$g \mbox{ s.t. } g(l) \in \exp \left( o(l) \right) ,$$ we have the following bound on the expected Phase 2 regret per timestep in a good epoch $l$: $$ E \left[r_{2,l} | G_l \right] \le 2\alpha_V^* n \cdot \exp\left(- \gamma l \right) , $$ where $\gamma$ is a constant which depends on $U$ and $z^*$. \end{lem} Proof: Recall that $\alpha_u^* = f \left( u^T z^* \right)$, where $$f \left( \beta \right) = \dfrac{1}{1+\exp\left(-\beta\right)}$$ is strictly increasing and continuous. Thus $f^{-1}$ is well defined, strictly increasing and continuous. Recall that $$ V = \left\{ v \in U : \alpha_v^* = \max_{u \in U} \alpha_u^* \right\} $$ is the set of equally best arms. Because $f \left(u_i^T z \right)$ is continuous in $z$ and defined over $\mathbb{R}^n$, it follows that there exists a neighborhood of $z^*$, denoted $A$, such that $$A = \left\{ z \in \mathbb{R}^n : \arg\max_{u \in U} \alpha_u(z) \in V \right\}.$$ Since $\Sigma$ is full rank, $A$ must contain an open parallelotope centered at $z^*$, $$B_{z^*} \left( \delta \right) = \left\{z \in \mathbb{R}^n : \left\Vert \Sigma^T z - \Sigma^T z^* \right\Vert _{\infty} < \delta \right\},$$ where $\delta > 0$ and is largest possible. An example of the problem parameters and the induced region $A$ is shown in Figure \ref{fig:three_phase_A}. Consider any $z \in B_{z^*} \left( \delta \right)$. By definition, $$|u^T z - u^T z^*| < \delta,\ \forall u \in \Sigma.$$ This is equivalent to $$\left| f^{-1} \left( \alpha_u(z) \right) - f^{-1} \left( \alpha_u^* \right) \right| < \delta,\ \forall u \in \Sigma.$$ Since $f^{-1}$ is continuous, this is equivalent to having a set of constants \begin{align*} \left\{ \underline{\alpha}_u, \overline{\alpha}_u \right\} _{u \in \Sigma} \mbox{ s.t. } \underline{\alpha}_u < \alpha_u(z) < \overline{\alpha}_u , \end{align*} where \begin{align*} & \underline{\alpha}_u = f \left( f^{-1} \left( \alpha_u^* \right) - \delta \right) \mbox{ and} \\ & \overline{\alpha}_u = f \left( f^{-1} \left( \alpha_u^* \right) + \delta \right) ,\ \forall u \in \Sigma . \end{align*} For a Phase 2 timestep during a good epoch $l$, the algorithm forms the empirical average rewards $$\hat{\alpha}_{u,l} \in (0, 1),\ \forall u \in \Sigma .$$ By the discussion above, \begin{align*} & \underline{\alpha}_u < \hat{\alpha}_{u,l} < \overline{\alpha}_u, \ \forall u \in \Sigma \\ \implies & \left. \hat{z}_l \in B_{z^*} (\delta) \subseteq A \right. \\ \implies & \left. C_t = \arg\max_{u \in U} \{u^T \hat{z}_l\} \in V \right. \end{align*} and we will have chosen one of the best arms, accumulating zero regret. Note that during epoch $l$, $\hat{\alpha}_{u,l}$ is a sum of $l$ i.i.d. $\mbox{Ber}\left(\alpha_{u}^{*}\right)$ random variables, $\forall u\in\Sigma$. By the Chernoff bound, \begin{align*} P\left(\hat{\alpha}_{u,l} < \underline{\alpha}_{u} | G_l \right) \le & \exp\left[-l \cdot D\left(\underline{\alpha}_{u}||\alpha_{u}^{*}\right)\right] \mbox{, and} \\ P\left(\hat{\alpha}_{u,l} > \overline{\alpha}_{u} | G_l \right) \le & \exp\left[-l \cdot D\left(\overline{\alpha}_{u}||\alpha_{u}^{*}\right)\right],\\forall u \in \Sigma, \end{align*} where $D\left(p||q\right)= p \cdot \log \dfrac{p}{q} + \left( 1-p \right) \cdot \log \dfrac{1-p}{1-q}$ is the K-L divergence between two Bernoulli distributions. Let $\gamma=\min _{u \in \Sigma} \min \left\{ D\left(\underline{\alpha}_{u}||\alpha_{u}^{*}\right) , D\left(\overline{\alpha}_{u}||\alpha_{u}^{*}\right)\right\}$. Note that from the definitions of $\underline{\alpha}_{i}$ and $\overline{\alpha}_{i}$, it follows that $$\underline{\alpha}_u < \alpha_u^* < \overline{\alpha}_u, \ \forall u \in \Sigma.$$ Since $D\left(p||q\right)=0\iff p=q$, we have that $\gamma>0$. By the union bound, $$P\left(\exists u\in\Sigma:\hat{\alpha}_{u,l}\notin\left(\underline{\alpha}_{u},\overline{\alpha}_{u}\right) | G_l \right) \le 2n \cdot \exp\left(-\gamma l\right).$$ Reviewing the chain of implications, we have \begin{align*} & P\left(\hat{z}_l \notin A | G_l \right) \\ \le & P\left(\hat{z}_l \notin B_{z^*}(\delta) | G_l \right) \\ = & P\left( \left\Vert \Sigma^{T} \hat{z}_{l} - \Sigma^{T} z^{*} \right\Vert_{\infty} > \delta | G_l \right) \\ = & P\left( \exists u \in \Sigma: \left| u^{T}\hat{z}_l - u^{T}z^{*} \right| > \delta | G_l \right) \\ = & P\left( \exists u \in \Sigma:\left| f^{-1}\left(\hat{\alpha}_{u,l}\right) - f^{-1}\left(\alpha_u^{*}\right) \right|>\delta | G_l \right)\\ = & P\left( \exists u \in \Sigma:\hat{\alpha}_{u,l} \notin \left( \underline{\alpha}_u ,\overline{\alpha}_u \right) | G_l \right)\\ \le & 2n \cdot \exp\left(- \gamma l \right). \end{align*} Then, we have a bound on the expected per-timestep regret $r_{2,l}$ during Phase 2 of epoch $l$: \begin{align*} E \left[ r_{2,l} | G_l \right] = & E\left[r_{2,l} | \hat{z}_l \in A, G_l \right] \cdot P\left(\hat{z}_l \in A | G_l \right) \\ & + E\left[r_{2,l} | \hat{z}_l \notin A | G_l \right] \cdot P\left(\hat{z}_l \notin A | G_l \right)\\ \le & 0\cdot P\left(\hat{z}_l \in A | G_l \right) + \alpha_V^* \cdot P\left(\hat{z}_l \notin A | G_l\right)\\ \le & 2\alpha_V^* n \cdot \exp\left(-\gamma l \right) . \end{align*} \begin{flushright} $\square$ \end{flushright} \begin{lem}\label{lemma finite total phase 2} For the Two-Phase Algorithm on finite arms, for a given choice of scheduling function $$g \mbox{ s.t. } g(l) \in \exp \left( o(l) \right) ,$$ we have the following bound on the expected total Phase 2 regret up to timestep $T$: $$ E[R_{2,T}] \le \alpha_V^* n \left( 2 k_2 + L \right) , $$ where $k_2$ is a constant which depends on $U$ and $z^{*}$. \end{lem} Proof: By Lemmas \ref{lemma bad epoch} and \ref{lemma finite per timestep phase 2}, we can upper-bound the expected total Phase 2 regret, \begin{align*} & E[R_{2,T}] \\ \le & \sum_{l=1}^{L} \left\{ \Big[ P(G_l) \cdot E[r_{2,l}|G_l] + P( \neg G_l) \cdot E[r_{2,l}| \neg G_l] \Big] \cdot g(l) \right\} \\ \le & \sum_{l=1}^{L} \left\{ \left[ P(G_l) \cdot 2\alpha_V^* n \cdot \exp\left(-\gamma l \right) + P( \neg G_l) \cdot \alpha_V^* \right] \cdot g(l) \right\} \\ \le & \sum_{l=1}^{L} \left\{ \left[ 2\alpha_V^* n \cdot \exp\left(- \gamma l \right) + 2 \alpha_V^* n \cdot \exp\left( -k_1 l \right) \right] \cdot g(l) \right\} \\ \le & 2\alpha_V^* n \sum_{l=1}^{L'} \left\{ \left[ \exp\left(- \gamma l \right) +\exp\left( -k_1 l \right) \right] \cdot g(l) \right\} \\ & + 2\alpha_V^* n \sum_{l=L'+1}^{L} \dfrac{1}{2} \\ \le & \alpha_V^* n \left\{ 2\sum_{l=1}^{L'} \left\{ \left[ \exp\left(- \gamma l \right) +\exp\left( -k_1 l \right) \right] \cdot g(l) \right\} + L \right\} , \end{align*} where $$L'=\max\left\{ l: \left[ \exp\left(- \gamma l \right) +\exp\left( -k_1 l \right) \right] \cdot g(l) > \dfrac{1}{2}\right\}$$ is a constant, independent of sample-path, that depends on $U$ and $z^{*}$ (and is therefore unknown to the algorithm). However, since we have assumed $g(l) \in \exp \left( o(l) \right)$, it follows that $$\lim_{l \rightarrow \infty} \left[ \exp\left(- \gamma l \right) +\exp\left( -k_1 l \right) \right] \cdot g(l) = 0 ,$$ and thus $L'$ is finite. Let $$k_2 = \sum_{l=1}^{L'} \left\{ \left[ \exp\left(- \gamma l \right) +\exp\left( -k_1 l \right) \right] \cdot g(l) \right\}, $$ which is well defined since $L'$ is finite. Thus, $$ E[R_{2,T}] \le \alpha_V^* n \left( 2 k_2 + L \right) . $$ \begin{flushright} $\square$ \end{flushright} \begin{thm}\label{theorem finite regret} For the Two-Phase Algorithm on finite arms, for a given choice of scheduling function $$g \mbox{ s.t. } g(l) \in \exp \left( o(l) \right) ,$$ we have the following bound on the expected total regret up to time-horizon $T$: $$ E \left[ R_{T} \right] \le 2 \alpha_V^* n \left( k_2 + g^{-1}(T) + 1 \right) . $$ \end{thm} Proof: Since the final epoch may be only partially finished, we will lower-bound the total time with the number of timesteps in the penultimate epoch's Phase 2, \begin{align*} T & \ge \sum_{l=1}^{L-1} \left\{ n + g(l) \right\} \ge g(L-1). \end{align*} Equivalently, $$L \le g^{-1}(T) + 1.$$ Then, using Lemmas \ref{lemma total phase 1} and \ref{lemma finite total phase 2}, \begin{align*} E[R_{T}] = & E[R_{1,T}] + E[R_{2,T}] \\ \le & 2 \alpha_V^* n \left( k_2 + L \right) \\ \le & 2 \alpha_V^* n \left( k_2 + g^{-1}(T) + 1 \right) . \end{align*} \begin{flushright} $\square$ \end{flushright} \begin{cor}\label{corollary finite regret asymptotic} For the Two-Phase Algorithm on finite arms, for a given choice of scheduling function $$g \mbox{ s.t. } g(l) \in \exp \left( o(l) \right) ,$$ we have the following asymptotic bound on the expected total regret up to time-horizon $T$: $$E \left[ R_{T} \right] \in O\left( n \cdot g^{-1}(T) \right).$$ \end{cor} Proof: By Theorem \ref{theorem finite regret}, as a function of $T$, \begin{align*} E[R_T] \le & 2 \alpha_V^* n \left( k_2 + g^{-1}(T) + 1 \right) \\ \in \ & O\left( n \cdot g^{-1}(T) \right), \end{align*} since $\alpha_V^* \le 1$, $k_2 > 0 $ is a constant dependent only upon $U$ and $z^*$, and $g^{-1}(T) \in \omega(1)$. \begin{flushright} $\square$ \end{flushright} \begin{lem}\label{lemma g g inverse asymptotics} $$g^{-1}(t)\in \omega \left( \log(t) \right) \implies g(l) \in \exp \left( o(l) \right).$$ \end{lem} Proof: \begin{align*} \lim_{t \rightarrow \infty} \dfrac{\log(t)}{ g^{-1}(t)} & = \lim_{l\rightarrow\infty} \dfrac{\log ( g(l) )}{ g^{-1}(g(l))} \tag{1} \\ & = \lim_{l\rightarrow\infty} \dfrac{\log(g(l))}{ l} \tag{2} \\ & = 0 \tag{3} , \end{align*} where (1) is by making the substitution $t = g(l)$, recalling that $g:\mathbb{N}_1 \rightarrow \mathbb{N}_0$ is strictly increasing by assumption, so $\left.\lim_{l \rightarrow \infty} g(l) = \infty \right.$. (2) is since by construction, $$g^{-1}(g(l)) = l, \ \forall l \in \mathbb{N}_1. $$ Lastly, (3) is since $g^{-1}(t)\in \omega \left( \log(t) \right)$, so by definition, $$\lim_{t \rightarrow \infty} \dfrac{g^{-1}(t)}{\log(t)} = \infty. $$ Hence $\log \left(g(l) \right)\in o(l)$, so $g(l) \in \exp \left(o(l) \right)$, and thus $g$ is a valid scheduling function. \begin{flushright} $\square$ \end{flushright} Let $\log^{*}(x)$, the iterated logarithm function, be defined recursively by $$ \log^{*}(x)= \begin{cases} 0, & \mbox{if }x\le 1\\ 1+\log^{*}\left(\log x\right), & \mbox{if } x>1 \end{cases}.$$ \begin{cor}\label{corollary gLLS} The Two-Phase Algorithm can achieve $\left. E[R_T] \in O \left( n \cdot \log(T) \cdot \log^{*} (T) \right) \right.$. \end{cor} Proof: Choose $$g_{LLS}(l) = \max \left\{ t \in \mathbb{N}_1 : \log(t) \cdot \log^*(t) \le l \right\}.$$ Then, $$g^{-1}_{LLS}(t) = \left\lfloor \log(t) \cdot \log^*(t) \right\rfloor,$$ $$\lim_{t \rightarrow \infty} \dfrac{g^{-1}_{LLS}(t)}{\log(t)} = \lim_{t \rightarrow \infty} \log^*(t) \rightarrow \infty .$$ Thus, $g_{LLS} \in \omega \left(\log(t) \right)$, and by Lemma \ref{lemma g g inverse asymptotics} and Corollary \ref{corollary finite regret asymptotic}, we have an achievable expected total regret of $$E[R_T] \in O \left( n \cdot g_{LLS}^{-1}(T) \right) \subseteq O \left( n \cdot \log(T) \cdot \log^{*} (T) \right).$$ \begin{flushright} $\square$ \end{flushright} \begin{remark} \label{Conjecture on Near-optimality} In accordance with other results, such as \cite{lai_robbins_85}, we suspect this problem has a lower bound that is asymptotically $c n \cdot \log(T)$, where $c$ is dependent on the problem parameters $U$ and $z^*$. If this is the case, then by including the term $\log^* (T)$, we are able to obtain an upper bound which is not tight, but within a factor of $\log^*(T)$, while avoiding a dependence on the problem parameters. \end{remark} \subsection{Generalization to Arm-dependent Rewards} Suppose that each arm $u\in U$ has a potentially different value of the reward, so that instead of a $\{0,1\}$ reward, it has a $\{0,w_u\}$ reward. Furthermore, suppose that $\{w_u\}_{u \in U}$ is known. Several definitions must be generalized, namely in the model, $$X_{t}\sim w_{C_{t}}\cdot\mbox{Ber}\left(\alpha_{C_{t}}^{*}\right) ,$$ $$ V = \left\{ v \in U : w_v \alpha_v^* = \max_{u \in U} w_u \alpha_u^* \right\} \subset U ,$$ $$ w_V \alpha_V^* = \max_{u \in U} w_u \alpha_u^* ,$$ $$A = \left\{ z \in \mathbb{R}^n : \arg\max_{u \in U} w_u \alpha_u(z) \in V \right\} ,$$ $$E[R_T] = E_{g}\left[\sum_{t=1}^{T} \left( w_u \alpha_V^* - X_{t} \right) \right] ,$$ and in the algorithm, $$ C_{(l)} \leftarrow \arg\max_{u \in U} w_u \alpha_u (\hat{z}_l) .$$ Then, Theorem \ref{theorem finite regret} generalizes with only minor modifications to the proof, yielding $$E[R_{T}] \le 2 w_V \alpha_V^* n \left( k_2 + g^{-1}(T) + 1 \right) .$$ Corollaries \ref{corollary finite regret asymptotic} and \ref{corollary gLLS} also generalize, with the same results as before. \end{cor}}

94 — 1111.4721

\caption{{\bf BIATECH-54 study design.} The log milligram (mg) fractional abundances (mg of protein/total mg in the sample) of proteins in Mix 1 and Mix 2 of the BIATECH-54 set. The \textcolor{blue}{blue}, \textcolor{green}{green} and \textcolor{red}{red} points represent 1, 2 and 6 proteins, respectively (54 total). Those proteins for which zero mg was included in either sample are shown on the axes.}

\caption{{\bf Schematic defining the distance between LC-MS features.} $d_{ijk}$ quantifies the time separation of the $i$'th yeast feature from the $j$'th UPS1 feature in sample $k$ (see Equation~\eqref{eq:dijk_definition}). The \textcolor{blue}{blue} and \textcolor{red}{red} lines represent the $2\sigma$ extent of yeast and UPS1 features, respectively.}

95 — 1111.4960

\caption{ {\itshape (Top panel) \swift/BAT lightcurve (in counts per second in the energy range 15-50 keV) and AGILE-GRID \gray flares for E$>$ 100 MeV as a function of time. The red arrows mark the dates of major \gray flares of \s: MJD 54572, 54772, 54811, 54998, 55001, 55003, 55007, 55025, 55034, 55324, 55343. Gray areas show the interval of good AGILE \gray exposure of \s. (Bottom panel) Average $\gamma$-ray flux from \s region in hard and soft X-ray state for E$>$100 MeV. } }

\caption{ {\itshape AGILE/GRID, AGILE/SA, Fermi/LAT, Swift/BAT, XTE/ASM, AMI-LA, RATAN-600 and Mets{\"a}hovi Radio Observatory data of \s during the uninterrupted 2-months period 2009 Jun.--Jul. (Panel 1 (top)) the AGILE-GRID \gray light curve with a variable window time to put in evidence the \cyg flares of 1-day timescale, the other data and upper-limits are determined with 1- 2- or 4-day time intervals; the Fermi light curve of 4-days timescale. E$>$100 MeV. (Panel 2) the AGILE-GRID \gray light curve for E$>$400 MeV and the coverage of MAGIC observation (gray area). (Panel 3) the hard X-ray light curve as monitored by BAT (15-50 keV) and by Super-AGILE (20-60 keV) with a daily timescale bin. (Panel 4) the soft X-ray light curve as monitored by XTE-ASM (1.3-12.1 keV) for a 1-day integration time bin; (Panel 5) the ASM hardness ratio $(5-12 keV)/(3-5 keV)$ data. (Panel 6) AMI-LA radio flux monitoring of \cyg at 15 GHz and the Mets{\"a}hovi Radio Observatory data at 37 GHz. (Panel 7) RATAN-600 radio telescope data at different frequencies. The AGILE/GRID $\gamma$-ray upper limits are at the $2-\sigma$ level, the flux error bars are $1-\sigma$ values. } }

\caption{ {\itshape Simultaneous AGILE/GRID, \swift/BAT and radio monitoring data of \s during the uninterrupted 18-months in the period 2009 Dec.--2011 May. (Top panel:) the AGILE-GRID \gray lightcurve for E$>$100 MeV. (Second panel:) \swift/BAT lightcurve (in counts per second in the energy range 15-50 keV). (Third panel:) AMI-LA radio flux monitoring of \cyg at 15 GHz.} }

96 — 1111.5286

\caption{\label{fig:2dFar} The critical point correlation functions for a gaussian random field in two dimensions, with Bessel function correlation $G\left(r\right)=J_0\left(\pi r\right)$. The term ${\cal C}_0$ represents the disconnect part of the correlation function, while ${\cal C}_1$ and ${\cal C}_2$ are the leading terms of the long range asymptotic expansion. The black dots represent the exact result obtained by numerical calculation explained in Appendix B.} \end{figure} \subsection{Three dimensions, $d=3$} \label{sec:three-dimensions} The critical-point density was found for $d=3$ in the context of the Cosmic Microwave Background radiation \cite{Vogel2008}. In this case there are six variables in the Hessian: $\boldsymbol{\psi} = (\phi_{xx},\phi_{xy},\phi_{xz},\phi_{yy},\phi_{yz},\phi_{zz})$, where as in the two-dimensional case, we write for brevity $\boldsymbol{\psi}=\boldsymbol{\psi}_1$ and the subscripts denote partial derivatives. Here we transform the Gaussian integral over $\boldsymbol{\psi}$ into an integral over the eigenvalues of the Hessian matrix $H$. To this end it is convenient to represent Gaussian exponent in the form \cite{Fyodorov2004,Vogel2008}: \begin{eqnarray*} \boldsymbol{\psi} \left[b(0)\right]^{-1}\boldsymbol{\psi} & = & \frac{1}{4\left\langle \phi_{xy}^{2}\right\rangle }\left(\Tr H^{2}-\frac{1}{d+2}{\Tr}^{2}H\right) \end{eqnarray*} where here we have used the block form of $\tilde{B}_0$ in Eq. (\ref{eq:B_Def}). Employing the standard approach of random matrix theory \cite{Mehta1991} one may transform from the integration variables $\boldsymbol{\psi}$ to to an integration over eigenvalues of $H$, $\boldsymbol{\lambda}=\left(\lambda_{1},\lambda_{2},\lambda_{3}\right)$. The measure of the integral in this case is \begin{eqnarray} d\psi \exp \left(\boldsymbol{\psi} \left[b(0)\right]^{-1}\boldsymbol{\psi} \right) = \label{eq:eigPDF} \\ \frac{ \prod_{j=1}^3 d\lambda_j}{ 8 \sqrt{2} \pi \langle \phi_{xy}^{2}\rangle^6} \left(\lambda_{2}-\lambda_{1}\right)\left(\lambda_{3}-\lambda_{2}\right)\left(\lambda_{3}-\lambda_{1}\right)\exp\left[-\frac{1}{4A^{2}}\left(\boldsymbol{\lambda} M\boldsymbol{\lambda}\right)\right] \nonumber, \end{eqnarray} with $M_{ij}=\delta_{ij}-\frac{1}{5}$. Following Ref. \cite{Vogel2008}, %\footnote{Note a slight numerical error in \cite{Vogel2008}, namely a factor $\frac{1}{2}$ in the exponent in Eq. 9 of that paper, as opposed to the correct factor $\frac{1}{4}$ as appearing in Eq. ~(\ref{eq:eigPDF})of this work, leads to different results from ours.%} further diagonalization of the Gaussian exponent followed by a change to spherical coordinates allows one to calculate $\langle\tilde{\rho}(0)\Tr H(0)\rangle$ and to deduce the value of the constant $f$ from Eq. (\ref{eq:fVal}). The result is $f=0$ for the correlation function of unsigned density, while for the minima-minima correlation: \begin{equation} \label{eq:fMin3} f = \frac{31250}{9 (29 - 6 \sqrt{6})^2 \pi} \end{equation} We conclude this section by referring to figure \ref{fig:3dFar}, which shows a comparison of our analytical asymptotic expression, and a numerical evaluation. \begin{figure}[h] \begin{centering} \subfloat[\emph{Minima-minima correlation function}]{\includegraphics[clip,width=0.8\textwidth]{figure2a}} \end{centering} \begin{centering} \subfloat[\emph{Unsigned correlation function}]{\includegraphics[clip,width=0.8\textwidth]{figure2b}} \end{centering} \caption{\label{fig:3dFar}The critical point correlation functions for a gaussian random field in three dimensions, with $G\left(r\right)=\frac{\sin\pi r}{\pi r}$. The meaning of lines and the dots in this figure is the same as in Fig.~1} \end{figure} \section{Discussion} \label{sec:discussion} In this work we calculated the asymptotic behavior of the pair correlation function of critical points, ${\cal C}(r)$ in two limiting regions, $r \to \infty$, and $r \to 0$. We have shown that in both limits the behavior correlation function depends on the types of densities of critical points, i.e. whether it is the unsigned density or the the minima-minima density. In the short range asymptotic it is manifested through the power law dependence of the correlation function: ${\cal C}(r) \propto r^{2-d}$ for of the unsigned correlation and ${\cal C}(r) \propto r^{5-d}$ for minima-minima correlations. In the long range asymptotics we have seen that there are two terms contributing to the correlation functions, see Eq.~(\ref{eq:main-result}). However, for the unsigned correlation the prefactor of the term proportional to $\nabla^4 G(r)$ vanishes while for the minima-minima correlation both terms contribute. Let us consider some specific examples in order to clarify the nature of the contributions of the two terms, $\nabla^4 G(r)$ and $\Tr \left[H_G(r)\right]^2$, to the long range asymptotic limit. Consider the case where the correlation function of the gaussian field is an oscillatory function of the form \begin{eqnarray} \lim_{r \to \infty} G(r) \longrightarrow A \frac{ \cos (r/\sigma)}{ r^\gamma}\label{eq:G1} \end{eqnarray} where $A$ is the amplitude, $\sigma$ is a constant which sets the length scale of the oscillations, while $\gamma >0$ determines the behavior of the envelope of the oscillations. This type of correlation describes, e.g., a field generated by wave chaos \cite{Dennis2007}. For this case in the asymptotic limit $ r \to \infty$ \begin{eqnarray} \nabla^4 G(r) \sim \frac{A}{\sigma^4}\frac{\cos (r/\sigma) }{r^\gamma} \label{eq:c1G1} \end{eqnarray} while \begin{eqnarray} \Tr \left[H_G(r)\right]^2 \sim \frac{A^2}{\sigma^4}\frac{\cos^2 (r/\sigma) }{r^{2\gamma}} \end{eqnarray} Thus for large enough $r$ the first term, Eq. (\ref{eq:c1G1}), dominate the behavior (unless its prefactor $\alpha_1$ vanishes as for the unsigned correlator). Consider the case where at large distances the correlation function is described by a simple power law behavior, i.e. \begin{eqnarray} \label{eq:G2} \lim_{r \to \infty} G(r) \longrightarrow \frac{ A }{ r^\gamma} \end{eqnarray} where $\gamma >-2$. This type of correlation, in particular for $\gamma < 0$, is believed to describe the potential landscape of glassy systems \cite{Fyodorov2007b}. For this case we find that \begin{eqnarray} \label{eq:c1G2} \nabla^4 G(r) \sim \gamma (2+ \gamma) (d-\gamma-2)(d-\gamma-4) \frac{ A}{r^{\gamma+4}}, \end{eqnarray} while \begin{eqnarray} \label{eq:c2G2} \Tr \left[H_G(r)\right]^2 \sim \gamma^2 (\gamma+ d)^2\frac{A^2}{r^{2\gamma+4}}. \end{eqnarray} Thus the term (\ref{eq:c1G2}) is dominant for $\gamma > 0$, while (\ref{eq:c2G2}) is dominant for $\gamma < 0$. Finally, for logarithmic correlation: \begin{eqnarray} \label{eq:G3} \lim_{r \to \infty} G(r) \longrightarrow A \log (r) \end{eqnarray} we have, \begin{eqnarray} \label{eq:d1G2} \nabla^4 G(r) \sim 2(2-d)(d-4) \frac{A}{r^{4}} \end{eqnarray} while \begin{eqnarray} \label{eq:d2G2} \Tr \left[H_G(r)\right]^2 \sim d \frac{A^2}{r^{4}} \end{eqnarray} These examples demonstrate that the asymptotic behavior of the correlations of critical points depend on the type of density, but are also sensitive to the particular value of $\gamma$ and the dimensionality $d$. For instance, for power law correlations with $\gamma=1$, the slower decaying term (\ref{eq:c1G2}) vanishes at $d=3$, while for the logarithmic correlation (\ref{eq:d1G2}) vanishes at $d=2$. \ack We would like to thank Y. Fyodorov for very useful discussions, and D. Klein for his advice on the numerical calculations in this work. This research has been supported by the United States-Israel Binational Science Foundation (BSF) grant No. 2008278. \appendix \section{Details of the derivation in Sec. 2} \label{sec:full-deriv-asympt} In this section we provide the details of the derivation of the results brought in Sec. \ref{sec:General-asymptotic-expressions}. To begin with, we present the form of the matrices introduced in Sec. \ref{sub:Presentation-of-the}, Eqs. (\ref{B-matrix}) and (\ref{eq:Correlation3}). As a first step, redefine the gaussian vector from Eq. (\ref{eq:Psi-def}): \begin{equation} \label{eq:psi-b-definition} \boldsymbol{\Psi} = \left(\boldsymbol{\nabla}\phi(\mathbf{r}_1), \boldsymbol{\nabla}\phi(\mathbf{r}_2), \boldsymbol{\psi}_1, \boldsymbol{\psi}_2 \right). \end{equation} This reordering of elements is more convenient in terms of the algebra to be carried out shortly. Then we have: \begin{eqnarray} \label{eq:b-full-def} B & = & \langle\boldsymbol{\Psi}\otimes\boldsymbol{\Psi}\rangle = \left(\begin{array}{cc} A & M \\ M^t & \beta \end{eqnarray}\end{array}\right)\\ & = & \left(\begin{array}{cc} \left(\begin{array}{cc} a(0) & a(r) \\ a(r) & a(0) \end{array}\right)& \left(\begin{array}{cc} 0 & m(r)\\ -m(r) & 0 \end{array}\right)\\\left(\begin{array}{cc} 0 & -m(r)^{t}\\ m(r)^{t} & 0 \end{array}\right)& \left(\begin{array}{cc} b(0) & b(r)\\ b(r) & b(0) \end{array}\right) \end{array}\right), \end{eqnarray} with the definitions \begin{eqnarray} a(r)_{ij} & = & -\left.\frac{\partial^2 G(\left|\mathbf{r}\right|)}{\partial r_i \partial r_j} \right|_{\mathbf{r}=(r,0\cdots,0)},\\ b(r)_{ij,kl} & = &\left.\frac{\partial^4 G(|\mathbf{r}|)}{\partial r_i \partial r_j \partial r_k \partial r_l} \right|_{\mathbf{r}=(r,0\cdots,0)},\\ m(r)_{i,jk} & = & \left.\frac{\partial^3 G(|\mathbf{r}|)}{\partial r_i \partial r_j \partial r_k} \right|_{\mathbf{r}=(r,0\cdots,0)}. \end{eqnarray} Here, $\mathbf{r} = \mathbf{r}_2 - \mathbf{r}_1$, and $b$ and $m$ have double indices to denote elements of $\boldsymbol{\psi}_i$. To make things clearer, we show some of these matrices explicitly for the three-dimensional case: \begin{eqnarray*} a(0) & = & I_d \left|G^{\left(2\right)}\left(0\right)\right|, \end{eqnarray*} where $I_d$ is the three dimensional identity matrix, and \begin{eqnarray*} b(0) & = & \left(\begin{array}{cccccc} 3 & 0 & 0 & 1 & 0 & 1\\ 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0\\ 1 & 0 & 0 & 3 & 0 & 1\\ 0 & 0 & 0 & 0 & 1 & 0 \\ 1 & 0 & 0 & 1 & 0 & 3 \end{eqnarray*}\end{array} \right)\frac{G^{\left(4\right)}\left(0\right)}{3}, \\ m & = & \left(\begin{array}{cccccc} p & 0 & 0 & q & 0 & q\\ 0 & q & 0 & 0 & 0 & 0\\ 0 & 0 & q & 0 & 0 & 0 \end{array} \right); \begin{array}{l} p = \left.\frac{\partial^{3}G(\mathbf{r})}{\partial r_1^3} \right|_{\mathbf{r}=(r,0 \cdots,0)}\\ q = \left.\frac{\partial^{3}G(\mathbf{r})}{\partial r_1 \partial r_2^2} \right|_{\mathbf{r}=(r,0\cdots,0)}\\ \end{array} \end{eqnarray*} It can also be verified that $a(r)$ and $b(r)$ have non-zero elements exactly for the same indices as $a(0)$ and $b(0)$. We plug these definitions into Eq. (\ref{eq:Correlation2}) and perform the integration over the field gradients and their Fourier conjugates. This leads to Eq. (\ref{eq:Correlation3}) with \begin{eqnarray} \label{eq:tildeB-def} \tilde{B} & = &\beta - D \\ D & = & M^t A^{-1} M = \left( \begin{array}{cc} d_1(r) & d_2(r) \\ d_2(r) & d_1(r) \\ \end{array}\right)\end{eqnarray} where \begin{eqnarray} d_i(r) & = & \left[a(0)^2-a(r)^2\right]^{-1}(-1)^{i-1}a(s) \label{eq:d_i}\\ s & = & \left\{\begin{array}{cc} 0 & i = 1 \\ r & i = 2 \end{eqnarray}\end{array}\right.\nonumber \end{eqnarray} \noindent \subsection*{Long range asymptotics} \label{sec:long-range-appen} In order to obtain the long range asymptotics to second order in $\eta$ (see Eq.~(\ref{eq:CorrelationExpansion})), we expand the prefactor and gaussian exponent in Eq. (\ref{eq:Correlation3}): \begin{eqnarray*} \fl \exp\left[-\frac{1}{2}\boldsymbol{\pi}\tilde{B}\boldsymbol{\pi}\right]&=& e^{-\frac{1}{2}\boldsymbol{\pi}\tilde{B}_0\boldsymbol{\pi}}\left[1-\eta \cdot \frac{1}{2} \boldsymbol{\pi} \Delta\tilde{B}\boldsymbol{\pi}+\frac{1}{2}\eta^2\left(\frac{1}{2}\boldsymbol{\pi} \Delta\tilde{B}\boldsymbol{\pi}\right)^{2} - \eta^2 \frac{1}{2} \boldsymbol{\pi} D \boldsymbol{\pi}+\ldots\right]\\\fl \det A^{-1/2} & = & \left\{\left(\det a(0)\right)^{2}\cdot\det\left[I-\left(a(0)^{-1}a(r)\right)^{2}\right]\right\} ^{-1/2}\\ \fl & = &\frac{1}{\det a(0)}\cdot\left[1+\frac{1}{2}\tr \left(a(0)^{-1}a(r)\right)^{2}-\ldots\right], \end{eqnarray*} where here we have: \begin{eqnarray*} \Delta\tilde{B} & = & \left(\begin{array}{cc} 0 & b(r) \\ b(r) & 0 \end{eqnarray*}\end{array}\right)\\ D & \simeq & \left(\begin{array}{cc} d_1(r) & 0 \\ 0 & d_1(r) \end{array}\right)\end{eqnarray*} since at the long range limit $d_2 \ll d_1$, as follows from Eq. (\ref{eq:d_i}). Then we obtain ${\cal C}_1$ from Eq. (\ref{eq:c11}) and ${\cal C}_2$ from Eq. (\ref{eq:c12}), which we can now write in full: \begin{eqnarray} \label{eq:c2-full} \fl{\cal C}_2(r) = & &\frac{1}{2}\left[{\cal C}_0\Tr\left[a(0)^{-1} a(r) \right]^2 - {\cal C}_0 \Tr (\tilde{B}_0 L) + \langle\tilde{\rho}\rangle^{-1} \Tr (F L)\right]\nonumber\\ & + & \frac{1}{4}\left[{\cal C}_0 \Tr (\tilde{B}_0 K \tilde{B}_0 K) - 2 \Tr (F K \tilde{B}_0 K) + {\cal C}_0^{-1}\Tr (F K F K)\right]\end{eqnarray} with $L = \tilde{B}_0^{-1} D \tilde{B}_0^{-1}$ and $K,F$ as in Eqs. (\ref{eq:K}) and (\ref{eq:F}). {\color{\myCol} We note however that if $G(r)$ diverges at the limit $r\to\infty$, then the $\det A^{-1/2}$ prefactor must be expanded to order $\eta^4$ so as to get the proper subleading behaviour of $\mathcal{C}(r)$.} Finally, let us examine $F$. We can write it as: \begin{equation} \label{eq:f-def} F \propto \left( \begin{array}{cc} \langle\tilde{\rho}(\mathbf{r}_1)\boldsymbol{\psi}_1\boldsymbol{\psi}_1\rangle_{1} \langle\tilde{\rho}(\mathbf{r}_2)\rangle_2 & \langle\tilde{\rho}(\mathbf{r}_1)\boldsymbol{\psi}_1\rangle_{1} \langle\tilde{\rho}(\mathbf{r}_2)\boldsymbol{\psi}_2\rangle_2 \\ \langle\tilde{\rho}(\mathbf{r}_1)\boldsymbol{\psi}_1\rangle_{1} \langle\tilde{\rho}(\mathbf{r}_2)\boldsymbol{\psi}_2\rangle_2 & \langle\tilde{\rho}(\mathbf{r}_1)\rangle_{1} \langle\tilde{\rho}(\mathbf{r}_2)\boldsymbol{\psi}_2\boldsymbol{\psi}_2\rangle_2 \end{array} \right)\end{equation} where \begin{equation} \label{eq:b0-int} \langle\cdots\rangle_i = \frac{1}{\sqrt{\det\left[2\pi b(0)\right]}} \int d\psi_i \left(\cdots\right)exp\left[-\frac{1}{2}\boldsymbol{\psi}_i b(0)^{-1} \boldsymbol{\psi}_i\right]\end{equation} Now, recall that $\tilde{\rho} \propto \det H$. From symmetry properties of the determinant, one can verify that the only non-zero in $\langle\tilde{\rho}\boldsymbol{\psi}\rangle$ (the position $\mathbf{r}_i$ no longer matters) is \begin{equation*} \left\langle\tilde{\rho} \frac{\partial^2 \phi}{\partial r_i^2} \right\rangle= \frac{1}{d} \left\langle\tilde{\rho} \Tr H \right\rangle, \end{equation*} and the only non-zeros in $\langle\tilde{\rho}\boldsymbol{\psi}\boldsymbol{\psi}\rangle$ are ($i\neq j$): \begin{equation*} \left\langle\tilde{\rho} \left(\frac{\partial^2 \phi}{\partial r_i^2}\right)^2 \right\rangle, \left\langle\tilde{\rho} \frac{\partial^2 \phi}{\partial r_i^2} \frac{\partial^2 \phi}{\partial r_j^2} \right\rangle~\mbox{and}~ \left\langle\tilde{\rho} \left(\frac{\partial^2 \phi}{\partial r_i\partial r_j}\right)^2 \right\rangle. \end{equation*} \subsection*{Short range asymptotics} \label{sec:short-range-append} To obtain the short range asymptotics, we assume $G(r)$ is an even function of $r$ and expand it to sixth order: \[ G\left(r\right)=g_0-\frac{1}{2!}g_2 r^{2}+\frac{1}{4!}g_4 r^{4}-\frac{1}{6!}g_6 r^{6}+\ldots \] where $-g_2,g_4$, and $-g_6$ are respectively the second, fourth and sixth derivatives of $G\left(0\right)$. Next, we need to find the prefactor in Eq. (\ref{eq:Correlation3}) and transform $\tilde{B}$ to an appropriate form. To find the prefactor we expand $\det A$ at the limit $r \ll 1$. This yields: \begin{eqnarray*} \det A & = & \prod_{i=1}^{d}\left(a(0)_{ii}^{2}-a(r)_{ii}^{2}\right)+\cdots=\frac{\left(g_2g_4r^{2}\right)^{d}}{3^{d-1}}+\cdots \end{eqnarray*} After performing the rotation (\ref{eq:rotation}), an explicit calculation gives for three dimensions: \begin{eqnarray*} \fl \Delta b &=&\left(\begin{array}{cccccc} -\frac{g_4^{2}}{2g_2}+\frac{g_6}{2} & 0 & 0 & -\frac{g_4^{2}}{6g_2}+\frac{g_6}{10} & 0 & -\frac{g_4^{2}}{6g_2}+\frac{g_6}{10}\\ 0 & -\frac{g_4^{2}}{18g_2}+\frac{g_6}{10} & 0 & 0 & 0 & 0\\ 0 & 0 & -\frac{g_4^{2}}{18g_2}+\frac{g_6}{10} & 0 & 0 & 0\\ -\frac{g_4^{2}}{6g_2}+\frac{g_6}{10} & 0 & 0 & -\frac{g_4^{2}}{18 g_2}+\frac{g_6}{10} & 0 & -\frac{g_4^{2}}{18 g_2}+\frac{g_6}{10}\\ 0 & 0 & 0 & 0 & \frac{g_6}{10} & 0 \\ -\frac{g_4^{2}}{6g_2}+\frac{g_6}{10} & 0 & 0 & -\frac{g_4^{2}}{18g_2}+\frac{g_6}{10} & 0 & -\frac{g_4^{2}}{18g_2}+\frac{g_6}{10} \end{array}\right)~ r^{2}+\Or\left(r^{4}\right)\\ \fl b_{0} & = & \left(\begin{array}{cccccc} 0 & 0 & 0 & \frac{g_6r^{2}}{45} & 0 & \frac{g_6r^{2}}{45}\\ 0 & 0 &0 &0 &0 &0 \\ 0 & 0 & 0 &0 &0 &0 \\ \frac{g_6r^{2}}{45} & 0 & 0 & \frac{16g_4}{9}-\frac{2g_6r^{2}}{27} & 0 & \frac{4g_4}{9}-\frac{2g_6r^{2}}{27}\\ 0 & 0 & 0 & 0 & \frac{2g_4}{3}-\frac{g_6r^{2}}{10} & 0 \\ \frac{g_6r^{2}}{45} & 0 & 0 & \frac{4g_4}{9}-\frac{2g_6r^{2}}{27} & 0 & \frac{16g_4}{9}-\frac{2g_6r^{2}}{27} \end{array}\right)+\Or\left(r^{4}\right) \end{eqnarray*} where $b_0 = b_1 + \Delta b$. It is now useful to diagonalize $b_0$. Let us rewrite the vector $\boldsymbol{\psi}_{+}$, as defined in Eq.~(\ref{eq:rotation}), in the form \begin{equation*} \boldsymbol{\psi}_{+} = \boldsymbol{\psi}_{\parallel}+\boldsymbol{\psi}_{\perp}, \end{equation*} where \begin{eqnarray*} \boldsymbol{\psi}_{\parallel} & = & \left(\Phi_{11},\Phi_{12},\dots,\Phi_{1d},0,\dots\right),\\ \boldsymbol{\psi}_{\perp} & = & \left(0,\dots,\Phi_{22},\Phi_{23},\dots,\Phi_{dd}\right). \end{eqnarray*} Here $\boldsymbol{\psi}_{\parallel}$ holds the first $d$ elements of $\boldsymbol{\psi}_+$, or in other words all those elements involving derivatives along the line connecting $\mathbf{r}_1$ and $\mathbf{r}_2$. $\boldsymbol{\psi}_{\perp}$ holds all the other elements namely those involving derivatives in direction transverse to the vector $\mathbf{r}_2-\mathbf{r}_1$. We then write the cross correlation matrix of $\boldsymbol{\psi}_+$ in the form \begin{equation*} b_0 = \left(\begin{array}{cc} c_0 & c_1\\ c_1^{t} & c_2 \end{array}\right) \end{equation*} where $c_0 = \langle\boldsymbol{\psi}_{\parallel}\otimes\boldsymbol{\psi}_{\parallel}\rangle,c_1 = \langle\boldsymbol{\psi}_{\parallel}\otimes\boldsymbol{\psi}_{\perp}\rangle$, and $ c_2 = \langle\boldsymbol{\psi}_{\perp}\otimes\boldsymbol{\psi}_{\perp}\rangle$. This matrix can now be block-diagonalized: \begin{eqnarray*} b_0 & = & u^{t} \left( \begin{array}{cc} c_0-c_1c_2^{-1}c_1^{t} & 0\\ 0 & c_2 \end{array} \right) u = u^t \left(\begin{array}{cc} c_3 & 0\\ 0 & c_2 \end{array}\right)u \end{eqnarray*} with \begin{eqnarray*} u &= & \left( \begin{array}{cc} I & 0\\ c_1c_2^{-1} & I \end{array} \right) \end{eqnarray*} Although in principal this requires transforming $\boldsymbol{\psi}_{\parallel},\boldsymbol{\psi}_{\perp}$ to new coordinates, in practice, $c_0 =\Or(r^4), c_1=\Or(r^2), c_2=\Or(1)$ so that to fourth order in the $r$ expansion of $\mathcal{C}(r)$, $u$ can be taken as the identity matrix. Substituting these expressions for the vectors and matrices into Eq.~(\ref{eq:Correlation3}), performing the inverse Fourier transform, and rescaling the coordinates as: \begin{eqnarray*} \boldsymbol{\psi}_{\parallel}\to r^{2}\boldsymbol{\psi}_{\parallel} & , &~ c_3 \to r^{-4}c_3\\ \boldsymbol{\psi}_{-}\to r\boldsymbol{\psi}_{-} & , &~ \Delta b \to r^{-2}\Delta b \end{eqnarray*} converts $\tilde{B}$ into a block-diagonal form, and thus \begin{equation*} \boldsymbol{\psi}\tilde{B}^{-1}\boldsymbol{\psi} \to \boldsymbol{\psi}_{+} \left( \begin{array}{cc} c_3 & 0 \\ 0 & c_2 \end{array} \right)^{-1} \boldsymbol{\psi}_{+} + \boldsymbol{\psi}_{-} \Delta b^{-1} \boldsymbol{\psi}_{-} \end{equation*} This form is convenient for expansion in $r$, yet in order to evaluate the integral Eq. ~(\ref{eq:Correlation3}) one should also expand $\tilde{\rho}$ as a power series in $r$. Here care must be taken, as $\tilde{\rho}$ is not a smooth function. Let us define a shorthand notation: \begin{equation*} h(\boldsymbol{\psi}_i) = \det H[\phi(\mathbf{r}_i)]. \end{equation*} so that \begin{equation*} \tilde{\rho}(0)\tilde{\rho}(r) = h(\boldsymbol{\psi}_1)h(\boldsymbol{\psi}_2) Q(\boldsymbol{\psi}_1,\boldsymbol{\psi}_2) \end{equation*} where $Q$ is some function which depends on the type of the density we are dealing with (unsigned or minima). Then we can expand the Hessian determinants as \begin{eqnarray*} \fl h\left(\boldsymbol{\psi}_1\right)h\left(\boldsymbol{\psi}_2\right) & = & h\left(\frac{\boldsymbol{\psi}_{\perp}+r^{2}\boldsymbol{\psi}_{\parallel}+r\boldsymbol{\psi}_{-}}{\sqrt{2}}\right)h\left(\frac{\boldsymbol{\psi}_{\perp}+r^{2}\boldsymbol{\psi}_{\parallel}-r\boldsymbol{\psi}_{-}}{\sqrt{2}}\right)\\ \fl & = & h^{2}\left(\frac{\boldsymbol{\psi}_{\perp}}{\sqrt{2}}\right)+\frac{r^{2}}{2}\boldsymbol{\psi}_{-}\cdot\left[h\boldsymbol{\nabla} \otimes \boldsymbol{\nabla} h-\boldsymbol{\nabla} h\otimes\boldsymbol{\nabla} h \right]\left(\frac{\boldsymbol{\psi}_{\perp}}{\sqrt{2}}\right)\cdot\boldsymbol{\psi}_{-}+\Or\left(r^{4}\right) \end{eqnarray*} Now, recall that the Hessian matrix one can construct from the elements of $\boldsymbol{\psi}_{\perp}$ is \begin{eqnarray} \left(\begin{array}{ccc} 0 & 0 & 0\\ 0 & \Phi_{22} & \Phi_{23}\\ 0 & \Phi_{23} & \Phi_{33} \end{array}\right). \end{eqnarray} This implies that $h\left(\frac{\boldsymbol{\psi}_{\perp}}{\sqrt 2}\right)=0$ since both the first row and first column equal zero. In addition, of all the elements of the vector $\nabla h$ only the first element survives: \begin{eqnarray} \frac{\partial}{\partial_{\Phi_{11}}}\det{H}=\det H_{\perp}=\det\left(\begin{array}{cc} \Phi_{22} & \Phi_{23}\\ \Phi_{23} & \Phi_{33} \end{array}\right) \end{eqnarray}. This formula holds for any dimension. It allows us to write for the unsigned correlation: \begin{eqnarray}\label{eq:final-short-uns} \fl \left\langle \tilde{\rho}\left(0\right) \tilde{\rho}\left(r\right)\right\rangle & = & \left\langle \left|-\frac{r^{2}}{2}\boldsymbol{\psi}_{-}\left(\boldsymbol{\nabla}h \otimes\boldsymbol{\nabla}h\right) \boldsymbol{\psi}_{-}+\Or\left(r^{4}\right)\right|\right\rangle \nonumber \\ \fl & = & \frac{r^{2}}{2} \Delta b_{11}\left\langle \left(\det H_{\perp}\right)^{2}\right\rangle+\Or\left(r^{4}\right) \end{eqnarray} thus recovering our main result, Eq. (\ref{eq:main-result}), with: \begin{equation} \label{eq:alpha3} \alpha_3 = \frac{3^{\frac{d-1}{2}}}{\left(2\pi\sqrt{g_2 g_4}r\right)^{d}}\cdot \Delta b_{11}\left\langle \left(\det H_{\perp}\right)^{2}\right\rangle \end{equation} Note that $ \left\langle\left(\det H_{\perp}\right)^{2}\right\rangle$ is the expectation of a polynomial in gaussian variables and can be evaluated with Wick's theorem. For the minima-minima correlation we find: \begin{eqnarray*} \fl \left\langle\left[\tilde{\rho}\left(0\right)\tilde{\rho}\left(r\right)\right]\right\rangle & = & \left\langle\left[-\frac{r^{2}}{2}\boldsymbol{\psi}_{-}\left(\boldsymbol{\nabla}h \otimes\boldsymbol{\nabla} h\right)\boldsymbol{\psi}_{-}+\Or\left(r^{4}\right)\right]\right.\\&\times & \left.\Theta\left[-\frac{r^{2}}{2}\boldsymbol{\psi}_{-}(\boldsymbol{\nabla}h\otimes\boldsymbol{\nabla}h)\boldsymbol{\psi}_{-}+\Or\left(r^{4}\right)\right] \prod_{i=0}^{d-2}\Theta\left(\det{H_{\perp}}^{\left(i\right)}+\Or\left(r^2\right)\right)\right\rangle\nonumber\\ \fl & = & \Or\left(r^{4}\right)\end{eqnarray*} To obtain this result we have used the fact that $\boldsymbol{\psi}_{-}(\boldsymbol{\nabla}h \otimes\boldsymbol{\nabla}h) \boldsymbol{\psi}_{-}$ is negative semidefinite. Also, note the fact that all the minors of ${H}$ at both points can be approximated by ${H_{\perp}}$ so that the two sets of $\Theta$-functions (from the two densities) merge. Obtaining the next order is considerably harder. However, we can deduce its power law dependence from the following considerations. Let us concentrate on the minima-minima correlation, so that: \[ \left\langle \left[\tilde{\rho}\left(0\right)\tilde{\rho}\left(r\right)\right]\right\rangle \sim r^{2}\left\langle \left[-\chi^{2}h_{\perp}^{2}+\mu r^{2}\right]\Theta\left[-\chi^{2}h_{\perp}^{2}+\mu r^{2}\right]\right\rangle \] Here $\chi$ is the first element of $\boldsymbol{\psi}_{-}$, $h_{\perp}=\det H_{\perp}$, and $\mu$ is some function of $\boldsymbol{\psi}_{\perp}$, $\boldsymbol{\psi}_{\parallel}$, and $\boldsymbol{\psi}_{-}$. Now, for the $\Theta$ function to be nonzero we require \begin{equation} \label{eq:theta-condition} \frac{\chi^{2}\eta_{\perp}^{2}}{\mu}< r^{2}\ll1. \end{equation} We can essentially ignore $\mu$ in this equation since as long as $\mu$ is not of order $\sim1$ this condition doesn't hold, so we have $\chi^{2}\eta_{\perp}^{2}\ll r^{2}\ll1$. Let us change coordinates by replacing $\chi$ with $x=\chi\eta_{\perp}$, keeping all other variables. This gives: \begin{eqnarray} \left\langle\left[\tilde{\rho}\left(0\right)\tilde{\rho}\left(r\right)\right]\right\rangle & \sim & r^{2}\int dx\left\langle\left[-x^{2}+\mu r^{2}\right]\Theta\left[r^{2}-x^2\right]\right\rangle \nonumber \\ & = & r^{2}\left.\left(-\frac{x^{3}}{3}+\mu r^{2}x\right)\right|_{-r}^{r} \propto r^{5} \label{eq:final-short-min} \end{eqnarray} and therefore ${\cal C}\left(r\right) \propto r^{5-d}$ which is the result stated in Eq. (\ref{eq:main-result}) for $k=3$. For the unsigned correlation, there is no condition such as (\ref{eq:theta-condition}), so the integration will simply give a correction of order $r^{2}$ above the first order. This can be seen by noting the identity $\left|x\right|=2x~\Theta\left(x\right)-x$. \noindent \subsection*{Application of short order analysis in two and three dimensions} \label{sec:appl-short-append} Let us apply the short-distance result expressed in Eq. (\ref{eq:final-short-uns}) to the two and three dimensional problem. To this end we apply Wick's theorem to Eq.~(\ref{eq:alpha3}) and find that, in two dimensions, the unsigned correlation behaves as \begin{equation}{\cal C}\left(r\right)= \alpha_3 + \Or\left(r^2\right)\label{eq:CunsShort2d} \end{equation} with \begin{eqnarray} \alpha_3 = \frac{1}{6\sqrt{3}\pi^{2}} \left(k_{1}^2 - k_{0}^2\right), \end{eqnarray} where $k_{0}=g_4 / g_2$, and $k_{1}^2 = g_6 / g_2$, while for the minima-minima correlation we have ${\cal C}\left(r\right) \propto r^3 $. In the three dimensional case the short range asymptotics of the unsigned correlation function takes the form \begin{equation} \mathcal{C}(r) = \alpha_3 \frac{1}{r} + \Or\left(r\right)\label{eq:CunsShort3d} \end{equation} with \begin{eqnarray} \alpha_3 = \frac{29}{288\pi^{3}} \sqrt{k_0}\left(k_1^2 - k_0^2\right), \end{eqnarray} while for the minima-minima correlation we find $\mathcal{C}(r) \propto r^2$. To verify our analysis, we compare our results to numerical calculations. The correlation functions chosen are the same as those appearing in sections \ref{sec:app-two-dimensions} and \ref{sec:three-dimensions}. Figure \ref{fig:Short-range-correlation} shows the excellent agreement we obtained with numerical calculations. There is one problematic graph, namely \ref{fig:short2dAbs}. The reason for this is simply the technical difficulty in numerically extracting the next-leading order behavior for this problem. \begin{figure} \subfloat[\label{fig:short2dMin} Minima-minima correlation in 2D.]{\includegraphics[clip,width=0.45\textwidth]{figure3a} }\hfill{}\subfloat[{\label{fig:short2dAbs}Unsigned correlation in 2D.}]{\includegraphics[clip,width=0.45\textwidth]{figure3b} } \subfloat[\label{fig:short3dMin}Minima-minima correlation in 3D.]{\includegraphics[clip,width=0.45\textwidth]{figure3c} }\hfill{}\subfloat[{\label{fig:short3dAbs}Unsigned correlation in 3D.}]{\includegraphics[clip,width=0.45\textwidth]{figure3d} } \caption{The short range limit of correlation functions of critical points in two and three dimensions. Solid lines represent the analytic result expressed in Eq.~ (\ref{eq:main-result}) while dots represent the result of a numerical calculation (see Appendix B) which has been performed for a gaussian random field with correlations defined by Eq.~(\ref{eq:Gd}).}\label{fig:Short-range-correlation} \end{figure} \section{Details of numerical analysis} \label{sec:deta-numer-analys} In this section we outline the methods we used to evaluate correlations numerically. From Eq. (\ref{eq:Correlation3}) in Sec. \ref{sub:Presentation-of-the} it follows that one can represent the correlation in the form: \begin{eqnarray} \fl{\cal C}(r) = \frac{1}{(2\pi)^{d}\sqrt{\det A}}\cdot \frac{1}{(2\pi)^{2(n-d)}\sqrt{\det \tilde{B}}} \int d\psi \tilde{\rho} (0) \tilde{\rho(r)} \exp \left(-\frac{1}{2} \boldsymbol{\psi} \tilde{B}^{-1} \boldsymbol{\psi}\right)\end{eqnarray} Here the elements of both $A$ and $\tilde{B}$ are functions of $r$. Evaluating the $\det A$ term numerically poses no difficulty. In order to evaluate the integral, we perform a Cholesky decomposition \cite{Press2007} numerically on $\tilde{B}$: \begin{eqnarray} \label{eq:cholesky} \tilde{B} = V^t V \\ \boldsymbol{\psi} \to V^{-1} \boldsymbol{\psi} \end{eqnarray} This transforms the integral to an integral over an appropriate set of standard normal variables: $\boldsymbol{\psi}_i \sim N(0,1)$. The integral can then be estimated by standard Monte-Carlo integration. Our integrations were performed with Sobol numbers of the appropriate dimension generated by Mathematica 7. The evaluations of the two-dimensional problem required $2^{23}$ lattice points for every data point appearing in figure \ref{fig:2dFar} and $2^{24}$ lattice points for every data point in figure \ref{fig:Short-range-correlation}. For the three dimensional problem, we used $2^{25}$ lattice points for data points in figure \ref{fig:3dFar}, and between $0.875$ and $ 1.125 \times 2^{26}$ points (depending on the specific dataset) for data points in figure \ref{fig:Short-range-correlation}. \bibliographystyle{unsrt} \bibliography{correlation} \end{document} }\end{eqnarray*}}\end{equation}}}}\end{eqnarray*}}

97 — 1111.5551

\caption{Powers for single locus alternative models. Power is calculated for each dataset with 100 replications total for the binary or continuous traits simulated under the single locus alternative model with or without covariate effect. The {\color{blue}$\times$} indicates the median of powers by the GLEAM and {\color{red} \textbullet }\;denotes the median of powers by the method based on Bayesian likelihood ratio. The whiskers on each bar represent the minimal and maximal powers respectively. The effect sizes of local ancestries are equal to the multiplication of effect size multiplier$c$ and the proportion of African ancestry population. \label{fig:singleAlt}}

98 — 1111.5580

\caption{Variability index (described in Section~\ref{sec:var_stats}) against \emph{Kepler} magnitude for the Raw, PDC and ARC \emph{Kepler} Q1 data. The solid line, in the same position on each graph, shows the photometric uncertainty (see Section~\ref{sec:var_frac}). {\Ared The removal of true stellar variability at medium levels by the PDC can be seen in the dearth of stars around $R_{\mbox{var}} = 10^{3} - 10^{4}$ ppm in the middle panel.}}

\caption{Division of high and low variability dwarf stars (red solid line) at twice the solar value (red dashed line). The solar level is calculated from the active Sun level (cyan dashed line) and the noise (magenta solid line), which itself is a combination of background (blue dashed) and photon noise (green dashed). {\Ared The orange dashed line marks the position of the solar line as determined by B10.}}

\caption{Fractions of stars more variable than once and twice the solar level, and fractions of periodic stars in the high and low variability groups (see Section~\ref{sec:per} for definition). {\Ared The random uncertainty for each of the measurements is \textless\0.5\%, and is negligible in comparison to difference introduced by the choice of dividing line, demonstrated by the 1 and 2 times the solar value comparisons listed in this table (see Section~\ref{sec:var_frac} for further discussion).}}

\caption{{\Ared Density plot of effective temperature against $R_{\mbox{var}}$, showing the decrease in variability with increasing temperature for all the selected dwarf stars (\emph{top}) and for those with \emph{Kepler} magnitude \textless\14 (\emph{bottom}). This illustrates that the dearth of low variability starts at low temperatures is not a result of the increased noise floor of the cool, faint stars.} The dashed line shows the solar variability level and the solid line is twice the solar level.}

\caption{The spatial distribution of low (\emph{left}) and high (\emph{right}) variability stars for ARC (\emph{top}) and PDC (\emph{bottom}). Each has been subject to a random selection of equal numbers of stars per magnitude bin, before a random selection of 6000 for each panel was selected. {\Ared The PDC panels show a slight tendency for the high variability sample to be more concentrated towards the galactic plane, whereas this effect is not significant in the ARC, suggesting it is more effective at removing contamination related systematics.}}

\caption{Histogram showing the galactic latitude distribution of low (solid line) and high (dashed line) variability stars for each spectral type. Sample selection ensured the orientation of the \emph{Kepler} field on the plane did not introduce biases. An equal sample of high and low variability stars from each magnitude bin was also selected. {\Ared The increased number of high variability F stars at low galactic latitudes may arise from giant contamination of the sample (see Section~\ref{sec:stelprop}).}}

\caption{Period distribution for each spectral type, for stars where a significant period between 1hr and 16d has been detected (see Section~\ref{sec:per} for method and caveats), {\Ared showing an increase in period towards later spectral types.}}

99 — 1111.5611

\caption{Layer-dependent thermal conductivity of few-layer graphene. The 1-layer nanoribbons refer to graphene and the 5% \symbol{126}8 layers nanoribbons are similar to utra-thin graphite. The inset is the layer-dependent thermal conductivity for various width zigzag graphene. The average temperature is 325K.}

100 — 1111.6236

\caption{\label{fig3} (Multimedia online) Snapshots of H$_2^+$ electron-nuclear probability density distribution taken at $t=30 \ \mathrm{fs}$ for four cases, in which the laser parameters for (a) (\textcolor[rgb]{0.00,0.00,1.00}{Media 1}), (b) (\textcolor[rgb]{0.00,0.00,1.00}{Media 2}), (c) (\textcolor[rgb]{0.00,0.00,1.00}{Media 3}) and (d) (\textcolor[rgb]{0.00,0.00,1.00}{Media 4}) are the same as those in Figs.~\ref{fig2}(a),~\ref{fig2}(d),~\ref{fig2}(i) and~\ref{fig2}(l), respectively, except that the CEP is fixed at $0.5\pi$. The color scale is logarithmic.}

101 — 1111.6545

\caption{\label{fig:bulkdos} Bulk density of states for $V=-1.9t$ (\dashed), $V=0$ (\chain), and $V=2t$ (\full) calculated in the middle of a 400-site chain ($\eta=0.04t$).}

\caption{\label{fig:bandwidth} Position $\epsilon_{\rm p}/2t$ of the peaks in the bulk density of states calculated with DDMRG in 200-site chains (\opencircle) and renormalized ``hopping term'' $t^{*}/t$ from~\eref{eq:hopping} (\full) as a function of the interaction parameter $V$.}

\caption{\label{fig:logbulkdos} Bulk density of states for $V=-\sqrt{2}t$ (\opencircle) and $V=2t$ (\opensquare) calculated with DDMRG in a 1600-site chain ($\eta=0.01t$). The solid lines show the power law~\eref{eq:LLdos} with the exact exponent $\alpha=\frac{1}{4}$ for both cases.}

\caption{\label{fig:edgedosrepulsive} Boundary LDOS $D(j=1,\epsilon)$ calculated with DDMRG on the first site of a 400-site chain ($\eta=0.04t$) for $V=0$ (\chain), $V=1t$ (\dashed), and $V=2t$ (\full).}

\caption{\label{fig:edgedosattractive} Boundary LDOS $D(j=1,\epsilon)$ calculated with DDMRG on the first site of a 200-site chain ($\eta=0.08t$) for $V=0$ (\chain), $V=-1t$ (\dashed), and $V=-1.9t$ (\full).}

102 — 1112.0093

\caption{(Color online) Theoretical cascade boundary given by the first-order cascade condition~\eqref{cascadecondition} (dashed green line) and numerically simulated mean cascade size $\rho$ for $\rho_0=10^{-3}$ (color coded) on a duplex network of $N=10^5$ with ER layers of equal mean degrees $z$. Inset: $\rho$ vs $z$ for threshold $R=0.18$ and different $\rho_0 = 5\times10^{-4}$ (red \textcolor{red}{$\triangle$}), $10^{-3}$ (blue \textcolor{blue}{$\bigcirc$}), $5\times10^{-3}$ ($\square$), obtained from simulations (symbols) and from Eq.~\eqref{rhoequation} (lines).}

103 — 1112.0287

\caption{(color online) IT-NCSM ground-state energies for \elem{He}{4} and \elem{O}{16} as function of $N_{\max}$ for the NN+3N-induced and the NN+3N-full Hamiltonians for a range of flow parameters: $\alpha=0.04\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.05\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.0625\,\text{fm}^4$ (\symboltriangle[FGGreen]), $0.08\,\text{fm}^4$ (\symbolbox[FGViolet]). Solid symbols correspond to the exact 3N interaction, open symbols to the NO2B approximation. Error bars indicate the uncertainties of the threshold extrapolations of the IT-NCSM. Data points are connected by straight lines to guide the eye, beyond the largest $N_{\max}$ an exponential extrapolation fitted to the last four data points is shown. }

\caption{(color online) Comparison of the ground-state energies of \elem{O}{16} obtained in IT-NCSM and CCSD including 3N interactions at the NO2B level for the NN+3N-induced and the NN+3N-full Hamiltonians with $\alpha=0.04\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.05\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.0625\,\text{fm}^4$ (\symboltriangle[FGGreen]), and $0.08\,\text{fm}^4$ (\symbolbox[FGViolet]).}

\caption{(color online) CCSD ground-state energies for \elem{O}{16} and \elem{O}{24} as function of $e_{\max}$ for the three types of Hamiltonians (see column headings) using the NO2B approximation for a range of flow parameters: $\alpha=0.04\,\text{fm}^4$ (\symbolcircle[FGBlue]), $0.05\,\text{fm}^4$ (\symboldiamond[FGRed]), $0.0625\,\text{fm}^4$ (\symboltriangle[FGGreen]), and $0.08\,\text{fm}^4$ (\symbolbox[FGViolet]). The filled symbols for the NN+3N-full Hamiltonian are for the standard chiral 3N interaction with cutoff 500 MeV, the open symbols for a modified 3N interaction with cutoff 400 MeV (see text).}

104 — 1112.0636

\caption{Main ion flux $\left(\Gamma_p\right)$ dependence on the background electron density gradient ($-R\grad n_e/n_e=R/L_{n_e}$). NL~GENE and fluid data with protons as main ions. Parameters are $q=1.4$, $s=0.8$, $\epsilon=r/R=0.14$, $R/L_{T_i, Z}=2.0$, $R/L_{T_e}=5.0$, and $\tau=T_e/T_i=1.0$. The fluid data was obtained for $k_\theta \rho_s = 0.2$. The fluxes are normalised to $v_{T,i}n_e\rho_i^2/R^2$. The error bars indicate an estimated uncertainty of one standard deviation.} \label{fig:main_ions} \end{figure} \clearpage \begin{figure}[\figplacing] \centering \subfloat[dependence of the impurity peaking factor $\left(PF_Z\right)$ on the background density gradient \label{fig:omn_TEM}] {\includegraphics[width=\figwidth]{figure_2a}} \phantomcaption{} \end{figure} \begin{figure}[\figplacing] \ContinuedFloat \centering \subfloat[dependence of the impurity diffusivity and convective velocity ($D_Z$ and $RV_Z$) on the background density gradient \label{fig:D_RV}]{\includegraphics[width=\figwidth]{figure_2b}} \phantomcaption{} \end{figure} \begin{figure}[\figplacing] \ContinuedFloat \centering \subfloat[scaling of real frequency ($\omega_r$) and growthrate ($\gamma$) with the background density gradient \label{fig:omn_TEM_eigens}]{\includegraphics[width=\figwidth]{figure_2c}} \caption{Scalings of the impurity peaking factor ($PF_Z=-RV_Z/D_Z$) with the background electron density gradient ($R/L_{n_e}$), with parameters as in figure~\ref{fig:main_ions}. QL and fluid data have been acquired using $k_\theta\rho_s=0.2$. Figure~\ref{fig:D_RV} shows the diffusivities and pinches corresponding to the NL~GENE impurity peaking factors ($PF_Z$) in figure~\ref{fig:omn_TEM}. $D_Z$ and $RV_Z$ are normalised to $v_{T,i}\rho_i^2/R$. The eigenvalues in figure~\ref{fig:omn_TEM_eigens} are from fluid and GENE simulations, and are normalised to $c_s/R$. The error bars indicate an estimated uncertainty of one standard deviation.} \label{fig:omn} \end{figure} \clearpage \begin{figure}[\figplacing] \centering \includegraphics[width=\figwidth]{figure_3} \caption{Scaling of the impurity peaking factor ($PF_Z=-RV_Z/D_Z$) with impurity charge $Z$, with parameters as in figure~\ref{fig:main_ions}; $k_\theta\rho_s=0.2$ was used in the QL and fluid simulations. The error bars indicate an estimated uncertainty of one standard deviation.} \label{fig:Z} \end{figure} \clearpage %\begin{figure}[\figplacing] % \centering % \includegraphics[width=\figwidth]{figures/kratios} % \caption[$-\grad n_Z \sim R/L_{n_Z}$ scaling of the ratio $k_y^Z/k_y^p$]{$-\grad n_Z \sim R/L_{n_Z}$ scaling of the ratio between mean and maximum wave numbers for impurities and main ions ($Z$ and $p$ respectively) $\left(k_{y,Z}/k_{y,p}\right)$. %NL GENE data of TEM turbulence with protons as main ions. %$R/L_{n_Z}=1.0,2.0,4.0,6.0,8.0,10.0,16.0$, $R/L_{n_p}=8.0$ %} % \label{fig:kratios} %\end{figure} %\clearpage % %\begin{figure}[\figplacing] % \centering % \includegraphics[width=\figwidth]{figures/kspectra_same} % \caption[$k_y$ spectra for $Z$ and $p$ with $R/L_{n_Z} = R/L_{n_p}$]{$k_y$ spectra for impurities and main ions ($Z$ and $p$ respectively) with $R/L_{n_Z} = R/L_{n_p} = 8.0$. %NL GENE data of TEM turbulence with protons as main ions.} % \label{fig:kspect_same_omn} %\end{figure} % %\begin{figure}[\figplacing] % \centering % \includegraphics[width=\figwidth]{figures/kspectra_diff} % \caption[$k_y$ spectra for $Z$ and $p$ with $R/L_{n_Z} \ll R/L_{n_p}$]{$k_y$ spectra for impurities and main ions ($Z$ and $p$ respectively) with $R/L_{n_Z} = 1.0$ and $R/L_{n_p} = 8.0$. %NL GENE data of TEM turbulence with protons as main ions.} % \label{fig:kspect_diff_omn} %\end{figure} %\clearpage \end{document} }

105 — 1112.1252

\caption{Atomic oxygen (measured by MBMS: \color{blue}$\medbullet$\color{black}, measured by TALIF: \color{blue}$\medcirc$\color{black}) and ozone (\color{red}$\blacksquare$\color{black}) density as a function of the distance from the jet in ambient helium (applied electrode voltage: 230\,V$_{RMS}$, gas flow: 1.4\,slm He with 0.6\,\% O$_2$). The lines reflect the densities derived by the Model 1 and 2. The TALIF measurement has been scaled down by a factor of 0.27. The results from MBMS and TALIF measurements are already published \cite{Ellerweg2010}.}

\caption{Atomic oxygen (\color{blue}$\medbullet$\color{black}) and ozone (\color{red}$\blacksquare$\color{black}) density as a function of the distance from the jet in ambient air (applied electrode voltage: 230\,V$_{RMS}$, gas flow: 1.4\,slm He with 0.6\,\% O$_2$). The solid lines reflect the densities derived by the model.}

106 — 1112.1476

\caption{{\it Left:} Gaussian kernel ($\sigma=0.2^\circ$) smoothed count map for the ROI about the LMC after subtraction of the background model for the energy range 200 MeV -- 20 GeV and for a pixel size of $0.1^\circ \times 0.1^\circ$. Overlaid is the N(\hi) contour of $1\times10^{21}$ H~cm$^{-2}$ of the LMC to indicate the extent and shape of the galaxy. The boxes show the locations of the 6 point sources (background blazars) that were included in the background model. {\it Right:} Differential average \gray{} emissivity spectrum for the LMC. Data for models \hone\(black dots) and \htwo\(red dots) (see text) and the local ISM emissivity (\cite{Abdo2009b} -- grey data points) are shown. %Error bars for \hone\ and \htwo\ include statistical and %systematic errors and the uncertainties in the total gas mass. Model lines: solid, predicted total \gray{} emissivity computed in the framework of a one-zone model for \htwo; long dashed, $\pi^0$-decay; short dashed, bremsstrahlung; dotted, IC. See \cite{LMCLat} for details.}

107 — 1112.2210

\caption{ \ifsuppl \includegraphics[scale=#2]{#1} \fi #3\label{#4} }

108 — 1112.2517

\caption{\label{fig:distance} Asymptotic jet velocity $V_j$ as a function of the distance $H$ between the laser spot and the free surface for different capillary diameters and energies. The triangles represent the data for the 200$\,\mu$m tube at $E=232\,\mu$J (\textcolor{ForestGreen}{$\blacktriangle$}) and $E=165\,\mu$J (\textcolor{ForestGreen}{$\vartriangle$}) and the circles show the results for the 500$\,\mu$m tube at $E=458\,\mu$J (\textcolor{blue}{$\Circle$}) and $E=305\,\mu$J (\textcolor{blue}{$\CIRCLE$}). The solid and dashed lines are showing a -1 power law. Typical error bars are shown for a few data points.}

\caption{\label{fig:energy_all} (a) Asymptotic jet velocity $V_j$ as a function of the energy absorbed by the liquid in the capillary tubes. The squares (\textcolor{red}{$\square$}), triangles (\textcolor{ForestGreen}{$\vartriangle$}) and circles (\textcolor{blue}{$\Circle$}) represent the data for capillary tubes with 50$\,\mu$m, 200$\,\mu$m, and 500$\,\mu$m inner diameter, respectively. Each data point is the result of at least three measurements. The lines are linear fits to the data. The data for the 50$\,\mu$m diameter tube are shown on an expanded scale in (b). %(b) Asymptotic jet velocities $V_j$ for 50 $\mu$m of supersonic microjets as a function of the energy absorbed by the liquid in the capillary tubes with 50$\,\mu$m inner diameter. Each data point is the result of at least five measurements. The linear dependence on the energy is seen to hold also for supersonic speeds.}

\caption{\label{FocusOffset} (a) Asymptotic jet velocity $V_j$ (\textcolor{ForestGreen}{$\Circle$}) as a function of the horizontal focus displacement of the laser for the 200$\,\mu$m tube. % with $E$ \textcolor{red}{XXXX} and $H$ \textcolor{red}{XXXX}. The black line shows the normalized vaporized liquid volume determined by geometrical optics approximation in both graphs (right axis). (b) Asymptotic jet velocity $V_j$ as a function of the vertical focus displacement of the laser. %for $E$ \textcolor{red}{XXXX} and $H$ \textcolor{red}{XXXX}. The inserts show the capillary and the laser focus (triangle) and the directions of $l_h$ and $l_v$ as seen from the capillary opening.}

\caption{\label{fig:focus_variation_total} Asymptotic jet velocity $V_j$ as a function of the focus offset of the laser for different capillary diameters. The circles (\textcolor{blue}{$\Circle$}) represent measurements for the 500$\,\mu$m tube; the triangles (\textcolor{ForestGreen}{$\vartriangle$}) refer to 200$\,\mu$m tube; the squares (\textcolor{red}{$\square$}) are for the 50$\,\mu$m. Each data point is the result of at least three measurements.}

109 — 1112.4721

\caption{\label{fig:mean-field} Numerically exact time-averaged relative population imbalance $\bar{z}$ (*) as a function of the scaled interaction strength $\Lambda=U(N-1)/J$ compared to the approximation (\ref{zbar}) (\textcolor{blau}{-}). }

\caption{\label{fig:N} Numerically exact time-averaged relative population imbalance $\bar{z}$ (*) as a function of the scaled interaction $\Lambda=U(N-1)/J$ compared to the approximation (\ref{zbar_app}) (\textcolor{blau}{-}) for different particle numbers $N$. }

110 — 1112.5006

\caption{Classification error for simulated data-sets consisting of $K=2$ clusters as a function of the phenotypic differentiation between the clusters. The variable plotted on the y-axis is the proportion of misclassified individuals (after correction for potential label switching issues). The variable plotted on the x-axis is the Hotelling T statistic and assesses the magnitude of the phenotypic differentiation. Our method: red triangles ({\color{red} \footnotesize $\boldsymbol \triangle$}), {\sc Mclust}: black circles ({$\boldsymbol \circ$}). }

111 — 1112.5319

\caption{ Selected normalized molecular transitions toward the observed cores. The scale is shown in the bottom right spectrum. The normalized intensity axis ranges from -0.33 to 1, while the velocity axis spans 20~\kms\centered at the\vlsr. {\it Rows}: individual cores, labeled on the left-hand side of the figure, ordered by its \Av\peak.{\it Columns}: molecular transition, ordered by molecular families, labeled on the top of the figure. The spectra have been divided by $[ \Av/100 \, {\rm mag}]$ to mimic the abundance, where the \Av\value is that at the respective core center\citep{roman10} given below the core name. Each molecular transition has been multiplied by a factor, given below its name, to fit in a common scale. {\it Colors}: used to highlight the distinctive emission of the different core groups. \textcolor{blue}{blue}: ubiquitous lines, \textcolor{ForestGreen}{green}: dense-medium molecular transitions, \textcolor{BurntOrange}{orange}: molecular transitions typical in oxo-sulfurated cores (see Sect.~\ref{disc}), \textcolor{BrickRed}{red}: molecular transitions typical in deuterated cores, and {\bf black}: mostly undetected species.\label{fig_spec}}

112 — 1112.5369

\caption{Comparison of the preliminary results for the ratio~$R$ (radiatively corrected) with the previous measurements (for $Q^2 < 2 \; \text{GeV}^2$): \textcolor{red}{\normalsize $\lozenge$}~\cite{Yount(1962)}; \textcolor{red}{$\square$}~\cite{Browman(1965)}, first experiment; \textcolor{red}{\large $\triangle$}~\cite{Browman(1965)}, second experiment; \textcolor{red}{\large $\blacktriangledown$}~\cite{Anderson(1966)}; \textcolor{red}{\large $\times$}~\cite{Bartel(1967)}; \textcolor{red}{$\blacksquare$}~\cite{Bouquet(1968)}; \textcolor{red}{\large $\blacktriangle$}~\cite{Anderson(1968)}; \textcolor{red}{\Large $\circ$}~\cite{Mar(1968)}; {\Large \textbullet}~this experiment, run~I. The curves show the theoretical predictions~\cite{Blunden(2005), Blunden} for the ratio~$R$ due to the two-photon exchange. The solid (dashed) line corresponds to the kinematics of run~I (run~II).}