2014

1 — 1401.0003

\caption{ (Top left) Image of ALMA J010748.3$-$173028 in 880-$\micron$ continuum emission (contours) overlaid on \textcolor{red}{the} integrated intensity map of \textcolor{red}{the} emission line (background). Images are not corrected for primary-beam attenuation. The contours start at $5\sigma$ with a $5\sigma$ step, where $\sigma = 0.14$ mJy beam$^{-1}$ at the ALMA-J0107 position. Crosses mark positions of near-infrared peaks shown in Figure~2 of \citet{Imanishi07}. Partial large circle represents primary beam size at 99.75~GHz. Filled ellipses in bottom-left corner indicate synthesized beam sizes of emission line image (gray) and 880-$\micron$ images (black). (Top right) 100 GHz spectrum of ALMA J010748.3$-$173028 across the 1.875~GHz spectral window of ALMA Band~3. Flux densities are measured with a $4''$ aperture and corrected for primary beam attenuation ($0.79\times$). Inset shows closeup of the spectrum; solid curve is \textcolor{red}{the} best-fit Gaussian. (Bottom) $15'' \times 15''$ multiwavelength (millimeter to X-ray) images of ALMA J010748.3$-$173028. Contours show the ALMA 880-$\micron$ image and start at $4 \sigma$ with a separation of $4 \sigma$, where $\sigma = 0.5$~mJy beam$^{-1}$ is the noise level corrected for primary beam attenuation ($0.28\times$). Contours of SMA 1.3-mm image are drawn at $2\sigma$, $3\sigma$, and $4\sigma$, where $\sigma = 1.21$~mJy beam$^{-1}$. }

2 — 1401.0329

\caption[]{\small Scales recovered in various survey configurations, using the information provided by \url{https://science.nrao.edu/facilities/vla/docs/manuals/oss2014a/performance/resolution.} Our recommendations for probing the large scale diffuse cluster emission are highlighted in bold. Configurations and bands that lead to insufficient resolution or severely limited largest angular scales (LASs) recovered are denoted in {\color{red}red}. The confusion limit is provided by the VLA Exposure Calculator, currently available at \url{https://obs.vla.nrao.edu/expCalc/14A/evlaExpoCalc.jnlp}. We find that the confusion level that could be reached will be negligible for all but the deepest surveys (which are specifically designed to approach the confusion limit). We also note that complementary higher resolution information (e.g. S Band observations in B Configuration) are necessary for ICM weather (Section \ref{weather}), AGN feedback (Section \ref{agn}), and for constraining the flux contributions from compact sources when observing the large scale diffuse emission. We indicate this higher resolution requirement in italics in the table. }

3 — 1401.1623

\caption{\label{tab:lbo}Properties of \ac{lbo} crystal for \shg from $\SI{798}{nm}$ to \blue.}

\caption{Draw and picture of the cavity for the \shg of \blue with the \ac{lbo} crystal.}

4 — 1401.2473

\caption{\small \label{fig:wang} The rates of star formation as a function of time for four simulations of cluster formation \citep[adopted from][]{wang10a}. The curves from top left to bottom right are for models that include, respectively, neither turbulence nor magnetic field nor outflow feedback (line labelled \textit{None}), turbulence only (\textit{Turb}), turbulence and magnetic field (\textit{Turb+B}), and all three ingredients (\textit{Turb+B+outflow}, see text for discussion). \red{The dashed horizontal line indicates a star formation rate such that the depletion time $t_{\rm dep}$ is equal to the free-fall time $t_{\rm ff}$, while the dotted line indicates a star formation rate corresponding to $t_{\rm dep} = 10 t_{\rm ff}$.} }

\caption{\small \label{fig:krumholz09} Slices through a simulation of the formation of a 70 $\msun$ binary system, taken from \citet{krumholz09c}. Both panels show a region 6000 AU on a side; colors show volume density on a scale from $10^{-20} - 10^{-14}$ g cm$^{-3}$, and plus signs show stars. In the upper panel, arrows show the velocity, while in the lower panel they show the net radiation plus gravity force $\mathbf{f}_{\rm rad}+\mathbf{f}_{\rm grav}$, with the arrow direction indicating the force direction, and the arrow length scales by $|\mathbf{f}_{\rm rad}+\mathbf{f}_{\rm grav}|/|\mathbf{f}_{\rm grav}|$. \red{These slices show the simulation at a time of $1.0$ mean-density free-fall times, at which point the total stellar mass is $\approx 60$ $\msun$ and the mass of the primary is $\approx 36$ $M_\odot$.} }

5 — 1401.3657

\caption{\red{The field where the luminosity function of \citet{PZ1999} was obtained, overlaid on an optical image of M22. The luminosity function was obtained from only part of the WFPC2 field illustrated.}}

\caption{Convergence of the initial values of $N$ and $r_h$, where $N$ is the number of objects (single stars and binary stars) and $r_h$ is the half-mass radius, for the determination of initial conditions in the case of 100\% retention of black holes. The plot gives an impression of the range of values sampled by the code. \red{Large symbols give the first 20 iterates, medium-sized symbols give iterates 21--60, and small symbols give the remainder (101 iterations altogether).}}

\caption{Marginal dependence of $Z$ (a measure of goodness of fit) on the initial value of $N$ (the initial number of objects), for the determination of initial conditions in the case of 100\% retention of black holes. Default large values of $Z$ may occur if the model failed to reach the required age of 12Gyr. Large values may also occur close to the best-fitting value of $N$ (about $7.8\times10^5$) if other parameters are far from optimal. \red{The meaning of the symbols is given in the caption to Fig.\ref{fig:convergence}. The inset gives the evolution of the goodness-of-fit parameter with iteration number.}}

\caption{The surface brightness distribution of the Monte Carlo model (made with the new version of the code called MOCCA), with initial conditions in Column 5 of Table \ref{tab:ics}, compared with data from \citet{Tr1995}. The solid line is their Chebyshev fit to the observation data. \red{The result for the Monte Carlo model is obtained by treating each star as a spherical shell, and projecting the model on the sky. When a line of sight passes just inside the shell of a bright star the inferred surface brightness shows a spike, as in this model at a projected radius of about 15 arcsec.} }

\caption{The spatial distribution of the black holes and black hole binaries \red{in Model C} at 12Gyr, compared with the spatial distribution of all stellar mass (including black holes). No axis labels or ticks are given for the number of black holes binaries (i.e. binaries in which at least one component is a black hole); the curve represents the cumulative fraction of the number, scaled to the vertical size of the frame. \red{For comparison, the two black holes discovered by \citet{St2012} lie at {\sl projected} radii of 0.25 and 0.4pc, respectively.}}

6 — 1401.3706

\caption{\label{fig:EELS}\textbf{Co concentration dependence of Fe $L_{2,3}$ edges EELS} for \bfa\for$x = $ 0, 0.04, 0.06, and 0.08. The inset: The measured L$_3$/L$_2$ white line ratio (circles) changes with $6+x$, the count of $3d$ electrons assuming Co$^{3+}$, considerably more than \red{the benchmark} (line, see the text), indicating that Co substitution introduces charge carriers.}

7 — 1401.3816

\caption{PBLM example for the transient startup of harmonic motion. In all figures \textcolor[gray]{0.5}{\sL} is the true system function $f$ and $\bullet$ are the $n$=15 randomly sampled values of $f$ with Gaussian noise of $\sigma$=10\%. In figure (a), \GLML is the optimized Gaussian regularization network GLM fit. In figure (b), the true function $f$ is compared to \IML the simple IM $\Intr$ \eqref{eq:toyIM}, and \textcolor{magenta}{\dotL} the predictions of the additive model $f_{\cal A}$ \eqref{eq: line add}. In figure (c) \PBLML is the prediction of PBLM \eqref{eq: phase basis} using the phase-shifted basis $\p$. The PBLM results are more than twice as accurate than either the GLM or the additive model, despite the small training sample and the fact that the IM has 32\% error. }

\caption{Convergence study on the dependence of the test RMSE (a) and eDOF (b) on the number of training points for the GLM (\textcolor{blue}{$\bullet$}) and PBLM (\textcolor{red}{$\blacktriangle$}). The RMSE is scaled by the maximum response $\eta U^2/g=0.187$ and the potential flow RMSE is shown for reference (\IML). The training sets were selected using systematic reduction, and in all cases the $Fr$=0.266,0.32 data have been held out. Note that the PBLM is essentially converged using 12 training data points (3 points per speed), and at this level outperforms the GLM using 84 points.}

8 — 1401.4067

\caption{Main waves of the $\eta'(\to \pi^-\pi^+\gamma\gamma)\pi^-$ data. In \textcolor{red}{red}: the $\eta\pi^-$ data multiplied by the mass-dependent phase-space factor from Eq.~1.1, taking into account final-state branching fractions.}

\caption{Relative phases in $\eta'\pi^-$ and $\eta\pi^-$ (\textcolor{red}{red}) for selected partial waves.}

9 — 1401.4201

\caption{ The Hilbert spectrum of dissolved oxygen ($\ocircle$) and temperature ($\square$), in which the vertical line indicates the annual cycle. Power law behavior is observed for both curves on the range $0.01<\omega<0.5$ day$^{-1}$ and $2<\omega<20$ day$^{-1}$, corresponding to a time scale $2<T<100$ days and $1.2<T<12$ hours, respectively. The corresponding scaling exponents are $1.93\pm0.05$ and $1.87\pm0.08$ for large scales, and $1.68\pm0.10$ and $1.35\pm0.10$ for small scales respectively. For comparison, the Fourier power spectrum \red{is} also shown for both dissolved oxygen ($\triangledown$) and temperature ($\triangle$). The inset shows the compensated spectrum $h(\omega)\omega^2C^{-1}$ ($E(f)f^2C^{-1}$) with fitted $C$ to emphasize the observed power laws.}

\caption{ a) The global cross correlation coefficient $R(\tau)$ between the dissolved oxygen and the temperature. Note that there is a strong annual cycle with a phase difference \red{of} 30 days. The global correlation coefficient is $R(0)=-0.37\pm0.10$. b) The measured cross correlation coefficient $R(0,t_w)$ with a sliding window $t_w=180$ days. It shows that the relation between dissolved oxygen and temperature \red{varies} from time to time, indicating a multi-scale property of such data sets. }

\caption{ a) The IMF modes with a 3-year mean period. b) The measured cross correlation $R(\tau)$. They are negatively correlated with a phase difference \red{of} around 2 months. The overall correlation coefficient is $R(0)=-0.68$. The mean period is determined by the distance between the two local extrema points ($\oplus$). }

\caption{ The measured TDIC for the annual cycle. It shows a transition from negative positive correlation between 2007 and 2008. The overall correlation coefficient is -0.60, see Fig.\,\ref{fig:annual}\,b. The hole is the$R$ \red{cannot} pass the $t-$test. The horizontal axis is the location of the \red{centre} of the sliding window.}

\caption{ The measured TDIC for 30-day mean period on the range May 2004 $\sim$ January 2008. There are more than half \red{a} years strong correlation patterns between January 2006 and June 2006. The overall correlation coefficient is 0.17. }

10 — 1401.4210

\caption{(Color online) Measured pdf of the $X=\ln(\epsilon)$ and local averaged $X_{\tau}=\ln(\epsilon_{\tau})$ for several time scales in both dissipative ($\tau/\tau_{\eta}<10$) and inertial ($\tau/\tau_{\eta}\ge10$) ranges. For comparison, the normal distribution is illustrated by a solid line. Graphically, except for values with $\vert X_{\tau}\vert>4\sigma$, the measured pdfs agree well with the lognormal distribution. Note that for display clarity, the measured pdf has been centered and vertically shifted by plotting $p(X_{\tau})=p_{\tau}(X_{\tau})/p_{\max}(\tau)$, in which $p_{\max}(\tau)=\max_{X_{\tau}}\left\{p_{\tau}(X_{\tau})\right\}$. }\label{fig:pdf} \end{figure} \begin{figure} \centering \includegraphics[width=0.65\linewidth,clip]{./Fig02} \caption{(Color online) Measured max value of pdfs, $p_{\max}(\tau)-p_{\max}(0)$, in which the inertial range $10\le \tau/\tau_{\eta}\le 100$ is illustrated by a dashed line. The upper inset shows the measured $p_{\max}(\tau)$, in which the solid line is the power-law fitting. The lower inset shows the compensated curve $\left(p_{\max}(\tau)-p_{\max}(0) \right)\tau^{-0.81}$. A power-law behavior is observed in the inertial range $10<\tau/\tau_{\eta}<100$ with a scaling exponent $\alpha=0.81\pm0.03$. The statistical error of $\alpha$ is the difference between the scaling exponent fitted on the first and second half of the scaling range (in log scale). }\label{fig:pmax} \end{figure} \begin{figure} \centering \includegraphics[width=0.65\linewidth,clip]{./Fig03} \caption{(Color online) Measured Fourier power spectrum of $\ln(\epsilon)$. A power law behavior is observed on the range $0.01<f\tau_{\eta}<0.06$ with a scaling exponent $1.06\pm0.13$. The inset shows the compensated curve with fitted parameters. }\label{fig:psd} \end{figure} \begin{figure} \centering \includegraphics[width=0.65\linewidth,clip]{./Fig04} \caption{(Color online) Measured variance $\sigma^2(\tau)$ ($\ocircle$) of $\ln(\epsilon_{\tau})$ ($\ocircle$). A log-law is observed with a scaling exponent $\beta=0.30\pm0.01$ on the range $10<\tau/\tau_{\eta}<100$. The inset shows the compensated curve with fitted parameters to emphasize the log-law. The statistical error of $\beta$ is estimated as in Fig.\,\ref{fig:pmax}. }\label{fig:variance} \end{figure} \begin{figure} \centering \includegraphics[width=0.8\linewidth,clip]{./Fig05a}\\ \includegraphics[width=0.8\linewidth,clip]{./Fig05b} \caption{(Color online) Measured autocorrelation function $\rho(\tau)$ for the logarithm of the energy dissipation rate $\epsilon$ ($\ocircle$) and pseudo-dissipation $\epsilon_T$ ($\square$). a) lin-lin plot, in which the dashed line and solid line are respectively exponential and logarithmic fitting. b) semilogx plot and c) semilogy plot. An exponential law is observed respectively on the range $3<\tau/\tau_{\eta}<15$ (resp. $0.019<\tau/T_{L}<0.097$) for $\epsilon$ and on the range $0<\tau/\tau_{\eta}<15$ (resp. $0<\tau/T_{L}<0.097$) for $\epsilon_T$. Log-law fitting is observed on the range $1<\tau/\tau_{\eta}<15$ (resp. $0.0065<\tau/T_{L}<0.097$) with a scaling exponent $\beta\rq{}=0.30\pm0.01$ for the full dissipation $\epsilon$, verifying Eq. \ref{eq:covariance}. }\label{fig:autocorrelation} \end{figure} \begin{figure} \centering \includegraphics[width=0.65\linewidth,clip]{./Fig06} \caption{(Color online) Measured $M_q(t/\tau)$ for $q=2,3$ and $4$, in which the inertial range $10\le \tau/\tau_{\eta}\le 100$ is indicated by a dashed vertical line. Power law behavior is observed on this inertial range for all moments we considered here. The scaling exponents $K_L(q)$ are then estimated on this inertial range. }\label{fig:Qmoment} \end{figure} \begin{figure} \centering \includegraphics[width=0.65\linewidth,clip]{./Fig07} \caption{(Color online) Measured $K_L(q)$ ($\ocircle$) and $q-\zeta_{L}(2q)$ ($\square$) provided by a Hilbert method. For comparison, the curve predicted by the lognormal model with an intermittent parameter $\mu=0.30$ (solid line) and the log-Poisson based multifractal model (dashed line) are also shown. The errorbar is the difference between the scaling exponent fitted on the first and second half of the inertial range (in log scale). }\label{fig:Kq} \end{figure} %%%%%%%%%%%%%%% The dataset considered here is composed by Lagrangian velocity trajectories in a homogeneous and isotropic turbulent flow obtained from a $2048^3$ DNS simulation with a Reynolds number $Re_{\lambda}=400$. We recall briefly some key parameters of this database. There are $\sim 2\cdot10^5$ fluid tracer trajectories, each composed by $N=4720$ time sampling saved every $0.1 \tau_{\eta}$ time units, in which $\tau_{\eta}$ is the Kolmogorov time scale. Hence, we can access time scales in the range $0.1<\tau/\tau_{\eta}<236$. The integral time scale $T_L$ is estimated by using the well-known Kolmogorov scaling relation as, i.e., $T_L/\tau_{\eta}=Re^{1/2}\simeq 155$, in which $Re=\frac{3}{20}Re_{\lambda}^2$ \citep{Pope2000}. The full energy dissipation rate $\epsilon(t)$ is retrieved from this database along the Lagrangian trajectories. Previously, an inertial range $0.01<\omega\tau_{\eta}<0.1$ (resp. $10<\tau/\tau_{\eta}<100$ or $0.065<\tau/T_L<0.65$) has been reported for this database by using a Hilbert-based methodology \citep{Huang2013PRE}. We therefore focus on this inertial range in the following analysis. The details of this database can been found in \cite{Benzi2009PRE}. Figure \ref{fig:pdf} shows the measured pdf $p(X)$ and $p_{\tau}(X_{\tau})$ for time scales in the dissipative ($\tau/\tau_{\eta}<10$) and inertial ($\tau/\tau_{\eta}>10$) ranges. For display clarity, the measured $p_{\tau}(X_{\tau})$ have been centered and vertically shifted by taking $p(X_{\tau})=p_{\tau}(X_{\tau})/p_{\max}(\tau)$, in which $p_{\max}(\tau)=\max_{X_{\tau}}\left\{p_{\tau}(X_{\tau})\right\}$. For comparison, the Gaussian distribution is illustrated as a solid line. Graphically, the measured pdf is slightly deviating from the Gaussian distribution when $\vert X\vert >4\sigma$. This confirms that the lognormal assumption for the energy dissipation rate approximately holds also in the Lagrangian frame at least for the central part of the pdf for $\vert X\vert<4\sigma$. Figure \ref{fig:pmax} shows the measured maximum value of the pdf $p_{\max}({\tau}) -p_{\max}(0)$, in which the inertial range $10<\tau/\tau_{\eta}<100$ is indicated by a dashed line. Here $p_{\max}(0)$ is for the original energy dissipation rate. A power-law behavior is observed on the inertial range, i.e., \begin{equation} p_{\max}({\tau}) -p_{\max}(0)\sim \tau^{\alpha} \end{equation} with a measured scaling exponent $\alpha=0.81\pm0.03$. Note that the statistical error of $\alpha$ is the difference between the scaling exponent fitted on the first and second half of the inertial range (in log scale). In the upper inset, we show the measured $p_{\max}(\tau)$, in which the solid line is the power-law fitting. In the lower inset we show the compensated curve $ \left(p_{\max}({\tau}) -p_{\max}(0)\right)\tau^{-0.81}$ to emphasize the observed power-law. A plateau is observed in the inertial range. To our knowledge, it is the first time that this pdf scaling relation is found. We have no interpretation presently for this relation and for the value of its exponent $\alpha$. Figure \ref{fig:psd} shows the measured Fourier power spectrum $X=\ln(\epsilon)$. A power-law behavior is observed in the range $0.01<f\tau_{\eta}<0.06$ with a scaling exponent $1.06\pm0.13$. The inset shows a compensated curve by using the fitted parameters to emphasize the observed power-law behavior. Figure \ref{fig:variance} shows the measured variance $\sigma^2(\tau)$ of $\ln(\epsilon_{\tau}(t))$ ($\ocircle$). A log-law is observed for $\sigma^2(\tau)$ respectively on the range $10<\tau/\tau_{\eta}<100$ with a scaling exponent $\beta=0.30\pm0.01$. The inset shows the corresponding compensated curve to emphasize the observed log-law. Eq.\,\ref{eq:sigma} is thus verified. Note that the $1/f$ type Fourier power spectrum is a consequence of a multiplicative cascade. Hence, this result is consistent with the logarithmic decay of the variance observed here. We also note that \cite{Pope1990PoF} have proposed an Ornstein-Uhlenbeck process for the dissipation field, which would predict a Lorentzian (or Cauchy) spectrum, i.e. a $f^{-2}$ decreasing. Here the observed $1/f$ spectrum does not support the Ornstein-Uhlenbeck proposal. Figure \ref{fig:autocorrelation} shows the measured autocorrelation function $\rho(\tau)$ for both the logarithm of the energy dissipation rate $\epsilon$ ($\ocircle$) and pseudo-dissipation $\epsilon_T=\nu \frac{\partial u_i}{\partial x_j}\frac{\partial u_i}{\partial x_j}$ ($\square$): a) lin-lin plot, b) semilogx plot and c) semilogy plot, respectively. The measured $\rho_{\epsilon}(\tau)$ and $\rho_{\epsilon_{T}}(\tau)$ cross zero at $\tau/\tau_{\eta}\simeq 23$. This is consistent with the observation in Ref. \cite{Pope1990Lagrangian}. We test the log-law, i.e., Eq.\,\ref{eq:covariance} first, see Fig.\,\ref{fig:autocorrelation}\,b). A log-law is observed for the dissipation$\epsilon$ on the range $1<\tau/\tau_{\eta}<15$ ($0.0065<\tau/T_{L}<0.097)$ with a scaling exponent $\beta\rq{}=0.30\pm0.01$. However, the log-law is less pronounced for the pseudo-dissipation. The log-law is illustrated by a thick solid line in Fig.\,\ref{fig:autocorrelation}. We note that \cite{Pope1990PoF} observed an exponential decay of $\rho(\tau)$. Figure \ref{fig:autocorrelation}\,c) shows the measured$\rho(\tau)$ in semilogy plot. An exponential law is observed respectively on the range $3<\tau/\tau_{\eta}<15$ ($0.019<\tau/T_{L}<0.097$) for $\epsilon$ and $0<\tau/\tau_{\eta}<15$ ($0<\tau/T_{L}<0.097$) for $\epsilon_T$. The exponential law is represented by a dashed line in Fig.\,\ref{fig:autocorrelation}. Moreover, the exponential law of pseudo dissipation is more pronounced than the one of full dissipation. Visually, it is difficult to make a distinction between logarithmic and exponential laws. However, we note that an exponential decay as found by \cite{Pope1990PoF} is not compatible with the intermittency framework for Lagrangian statistics, which is now well accepted \citep{Chevillard2003PRL,Biferale2004PRL}. Despite the scaling range, Eq.\,\ref{eq:covariance} is verified. Note that the autocorrelation function can be related to the Fourier power spectrum via $\rho(\tau)=\int_0^{+\infty} E(f)\cos(2\pi f\tau)\upd f$, in which $E(f)$ is the Fourier power spectrum of $X$. Therefore, except for $f=(n+1/2)/2\tau$, $n=0,1,2,\cdots$, all Fourier modes contribute to $\rho(\tau)$, indicating a mixing of large- and small-scale information. This could be one reason for the shift of the scaling range. A similar phenomenon is observed for the structure-function, which could be understood as a finite size effect of the range of the power-law, known as infrared effect (large-scale motions) and ultraviolet effect (small-scale motions) \citep{Huang2010PRE,Huang2013PRE}. The intermittency parameter $\mu=2-\zeta_L(4)=0.30\pm0.14$ provided by the Hilbert method \citep{Huang2013PRE} is consistent with the scaling exponents $\beta$ and $\beta\rq{}$ we obtained here. Let us note here that the covariance log-relation was not an hypothesis of Kolmogorov, and is a relation which is different from Eq. \ref{eq:sigma}. However there are some relations between them: for a lognormal multiplicative cascade with intermittency parameter $\mu=K_L(2)$, it can be shown that the covariance of $X$ should have a log-law with parameter $\beta\rq{}=\mu$ \citep{Kahane1985}. Kolmogorov's hypothesis for the variance of $X_{\tau}$ is also a consequence of the cascade and its parameter is $\beta=\mu$. Here we find $\mu=0.30$ and the values for the slopes of the covariance and variance rescaling, are fully compatible with this value of the intermittency parameter. We also note that the direct estimation of the intermittency parameter from the Eulerian structure function is $\mu_E=2-\zeta(6)=0.34\pm0.03$ (not shown here). It is also compatible with the value we obtain for Lagrangian fluctuations. Furthermore, when the covariance has a logarithmic decay, the Fourier power spectrum has a $-1$ scaling, also found here. All these results are consistent, and confirm that the dissipation in the Lagrangian frame can be described by a multiplicative cascade. Figure \ref{fig:Qmoment} shows the measured $M_q(\tau)$ for $q=2,3$ and $4$. Power-law behavior is observed for all moments on the inertial range $10<\tau/\tau_{\eta}<100$. The scaling exponent $K_L(q)$ is then estimated on this range. Figure \ref{fig:Kq} shows the measured $K_L(q)$ ($\ocircle$) on the range $0<q<4$. The errorbar is the difference between the scaling exponent fitted on the first and second half of the inertial range (in log scale). For comparison, the $K_L(q)=q-\zeta_{L}(2q)$ ($\square$) provided by the Hilbert-based methodology is also shown. We estimated $\zeta_L(q)$ up to $q=6$ by using the Hilbert-based method \citep{Huang2013PRE}. The corresponding $q$ for $K_L(q)$ is $3$. The definition of the errorbar is the same as the one for $K_L(q)$. The $K_L(q)$ provided by \citet{Biferale2004PRL} log-Poisson based multifractal model and by the lognormal model with the intermittency parameter $\mu=0.30$ are respectively shown as a dashed and solid line. For $q\le3$, all symbols collapse, showing the validity of the scaling relation of Eq.\ref{eq:LK62} predicted by the LRSH. For $q>3$, the measured $K_L(q)$ deviates from the lognormal model since the high-order $M_q(\tau)$ corresponds the statistics of the tail of the pdf and we observed deviations from the Gaussian distribution when $X>4\sigma$. \section{Conclusion} In summary, the scaling statistics of the energy dissipation along the Lagrangian trajectory is investigated by using fluid tracer particles obtained from a high resolution direct numerical simulation with $Re_{\lambda}=400$. Both the energy dissipation rate $\epsilon$ and the local time averaged $\epsilon_{\tau}$ agree reasonably with the lognormal distribution hypothesis. The measured $p_{\max}(\tau)-p_{\max}(0)$ (maximum value of a pdf) obeys a power law with a scaling exponent $0.81$, a result for which we have no theoretical explanation. Several statistics of the energy dissipation are then examined. It is found that the autocorrelation function $\rho(\tau)$ of $\ln(\epsilon(t)) $ and variance $\sigma_{\tau}^2$ of $\ln(\epsilon_{\tau})$ obey log-laws with scaling exponents compatible with the intermittency parameter $\mu=0.30$ as expected for multiplicative cascades. These results show that the dissipation along Lagrangian trajectories can be modelled by multiplicative cascades. The $q$th-order moment of $\epsilon_{\tau}$ has a clear power-law on the inertial range. The LRSH assumptions Eqs.\,\ref{eq:sigma} and \ref{eq:moments}, and scaling relation \ref{eq:LK62} are then verified. \begin{acknowledgments} This work is sponsored by the National Natural Science Foundation of China under Grant (No. 11072139, 11032007, 11272196, 11202122 and 11332006) , \lq{}Pu Jiang\rq{} project of Shanghai (No. 12PJ1403500) and the Shanghai Program for Innovative Research Team in Universities. We thank Prof. F. Toschi for sharing his DNS database, which are freely available from the iCFD database and is available for download at {{http://cfd.cineca.it}}. \end{acknowledgments} \bibliographystyle{jfm} % Note the spaces between the initials %\bibliography{all} \begin{thebibliography}{32} \expandafter\ifx\csname natexlab\endcsname\relax\def{\}{natexlab}#1{#1}\fi \bibitem[Arneodo {\em et~al.\/}(1998)Arneodo, Bacry, Manneville \& Muzy]{Arneodo1998}{\sc Arneodo, A., Bacry, E., Manneville, S. \& Muzy, J.F.} 1998 {Analysis of Random Cascades Using Space-Scale Correlation Functions}. {\em Phys. Rev. Lett.\/} {\bf 80}~(4), 708--711. \bibitem[Benzi {\em et~al.\/}(2009)Benzi, Biferale, Calzavarini, Lohse \& Toschi]{Benzi2009PRE}{\sc Benzi, R., Biferale, L., Calzavarini, E., Lohse, D. \& Toschi, F.} 2009 Velocity-gradient statistics along particle trajectories in turbulent flows: The refined similarity hypothesis in the lagrangian frame. {\em Phys. Rev. E\/} {\bf 80}~(6), 066318. \bibitem[Biferale {\em et~al.\/}(2004)Biferale, Boffetta, Celani, Devenish, Lanotte \& Toschi]{Biferale2004PRL}{\sc Biferale, L., Boffetta, G., Celani, A., Devenish, B.J., Lanotte, A. \& Toschi, F.} 2004 Multifractal statistics of lagrangian velocity and acceleration in turbulence. {\em Phys. Rev. Lett.\/} {\bf 93}~(6), 064502. \bibitem[Borgas(1993)]{Borgas1993Lagrangian}{\sc Borgas, M.S.} 1993 The multifractal lagrangian nature of turbulence. {\em Phil. Trans. R. Soc. A\/} {\bf 342}~(1665), 379--411. \bibitem[Chen {\em et~al.\/}(1997)Chen, Sreenivasan, Nelkin \& Cao]{Chen1997PRL}{\sc Chen, S.Y., Sreenivasan, K.R., Nelkin, M \& Cao, N.Z.} 1997 Refined similarity hypothesis for transverse structure functions in fluid turbulence. {\em Phys. Rev. Lett.\/} {\bf 79}~(12), 2253--2256. \bibitem[Chevillard \& Meneveau(2006)]{Chevillard2006PRL}{\sc Chevillard, L. \& Meneveau, C.} 2006 Lagrangian dynamics and statistical geometric structure of turbulence. {\em Phys. Rev. Lett.\/} {\bf 97}~(17), 174501. \bibitem[Chevillard {\em et~al.\/}(2003)Chevillard, Roux, L{\'e}v{\^e}que, Mordant, Pinton \& Arn{\'e}odo]{Chevillard2003PRL}{\sc Chevillard, L., Roux, S.G., L{\'e}v{\^e}que, E., Mordant, N., Pinton, J-F \& Arn{\'e}odo, A.} 2003 Lagrangian velocity statistics in turbulent flows: Effects of dissipation. {\em Phys. Rev. Lett.\/} {\bf 91}~(21), 214502. \bibitem[Falkovich {\em et~al.\/}(2012)Falkovich, Xu, Pumir, Bodenschatz, Biferale, Boffetta, Lanotte \& Toschi]{Falkovich2012PoF}{\sc Falkovich, G., Xu, H.T., Pumir, A., Bodenschatz, E., Biferale, L., Boffetta, G., Lanotte, A.S. \& Toschi, F.} 2012 On lagrangian single-particle statistics. {\em Phys. Fluids\/} {\bf 24}~(4), 055102. \bibitem[Frisch(1995)]{Frisch1995}{\sc Frisch, U.} 1995 {\em{Turbulence: the legacy of AN Kolmogorov}\/}. Cambridge University Press. \bibitem[Homann {\em et~al.\/}(2011)Homann, Schulz \& Grauer]{Homann2011PoF}{\sc Homann, H., Schulz, D.L. \& Grauer, R.} 2011 Conditional eulerian and lagrangian velocity increment statistics of fully developed turbulent flow. {\em Phys. Fluids\/} {\bf 23}, 055102. \bibitem[Huang {\em et~al.\/}(2013)Huang, Biferale, Calzavarini, Sun \& Toschi]{Huang2013PRE}{\sc Huang, Y.X., Biferale, L., Calzavarini, E., Sun, C. \& Toschi, F.} 2013 Lagrangian single particle turbulent statistics through the hilbert-huang transforms. {\em Phys. Rev. E\/} {\bf 87}, 041003(R). \bibitem[Huang {\em et~al.\/}(2010)Huang, Schmitt, Lu, Fougairolles, Gagne \& Liu]{Huang2010PRE}{\sc Huang, Y.X., Schmitt, F.G., Lu, Z.M., Fougairolles, P., Gagne, Y. \& Liu, Y.L.} 2010 {Second-order structure function in fully developed turbulence}. {\em Phys. Rev. E\/} {\bf 82}~(2), 026319. \bibitem[Kahane(1985)]{Kahane1985}{\sc Kahane, J.P.} 1985 Sur le chaos multiplicatif. {\em Ann. Sci. Math. Que.\/} {\bf 9(2)}, 105--150. \bibitem[Kholmyansky \& Tsinober(2009)]{Kholmyansky2009PLA}{\sc Kholmyansky, M. \& Tsinober, A.} 2009 On an alternative explanation of anomalous scaling and how well-defined is the concept of inertial range. {\em Phys. Lett. A\/} {\bf 373}~(27), 2364--2367. \bibitem[Kolmogorov(1962)]{Kolmogorov1962}{\sc Kolmogorov, A.N.} 1962 {A refinement of previous hypotheses concerning the local structure of turbulence in a viscous incompressible fluid at high Reynolds number}. {\em J. Fluid Mech.\/} {\bf 13}, 82--85. \bibitem[Meneveau(2011)]{Meneveau2011ARFM}{\sc Meneveau, C.} 2011 Lagrangian dynamics and models of the velocity gradient tensor in turbulent flows. {\em Annu. Rev. Fluid Mech.\/} {\bf 43}, 219--245. \bibitem[Mordant {\em et~al.\/}(2002)Mordant, Delour, L{\'e}veque, Arn{\'e}odo \& Pinton]{Mordant2002PRL}{\sc Mordant, N., Delour, J., L{\'e}veque, E., Arn{\'e}odo, A. \& Pinton, J.-F.} 2002 Long time correlations in lagrangian dynamics: a key to intermittency in turbulence. {\em Phys. Rev. Lett.\/} {\bf 89}~(25), 254502. \bibitem[Pope(1990)]{Pope1990Lagrangian}{\sc Pope, S.B.} 1990 Lagrangian microscales in turbulence. {\em Phil. Trans. R. Soc. A\/} {\bf 333}~(1631), 309--319. \bibitem[Pope(2000)]{Pope2000}{\sc Pope, S.B.} 2000 {\em{Turbulent Flows}\/}. Cambridge University Press. \bibitem[Pope \& Chen(1990)]{Pope1990PoF}{\sc Pope, S.B. \& Chen, Y.L.} 1990 The velocity-dissipation probability density function model for turbulent flows. {\em Phys. Fluids\/} {\bf 2}, 1437. \bibitem[Praskovsky {\em et~al.\/}(1997)Praskovsky, Praskovskaya \& Horst]{Praskovsky1997PoF}{\sc Praskovsky, A., Praskovskaya, E. \& Horst, T.} 1997 Further experimental support for the kolmogorov refined similarity hypothesis. {\em Phys. Fluids\/} {\bf 9}~(9), 2465--2467. \bibitem[Sawford \& Yeung(2011)]{Sawford2011PoF}{\sc Sawford, B.L. \& Yeung, P.K.} 2011 Kolmogorov similarity scaling for one-particle lagrangian statistics. {\em Phys. Fluids\/} {\bf 23}, 091704. \bibitem[Schmitt(2003)]{Schmitt2003}{\sc Schmitt, F.G.} 2003 {A causal multifractal stochastic equation and its statistical properties}. {\em The European Physical Journal B\/} {\bf 34}~(1), 85--98. \bibitem[Sreenivasan \& Antonia(1997)]{Sreenivasan1997}{\sc Sreenivasan, K.R. \& Antonia, R.A.} 1997 {The phenomenology of small-scale turbulence}. {\em Annu. Rev. Fluid Mech.\/} {\bf 29}, 435--472. \bibitem[Stolovitzky {\em et~al.\/}(1992)Stolovitzky, Kailasnath \& Sreenivasan]{Stolovitzky1992PRL}{\sc Stolovitzky, G., Kailasnath, P. \& Sreenivasan, K.R.} 1992 Kolmogorov's refined similarity hypotheses. {\em Phys. Rev. Lett.\/} {\bf 69}~(8), 1178. \bibitem[Stolovitzky \& Sreenivasan(1994)]{Stolovitzky1994}{\sc Stolovitzky, G. \& Sreenivasan, K.R.} 1994 {Kolmogorov's refined similarity hypotheses for turbulence and general stochastic processes}. {\em Rev. Mod. Phys.\/} {\bf 66}~(1), 229--240. \bibitem[Toschi \& Bodenschatz(2009)]{Toschi2009ARFM}{\sc Toschi, F. \& Bodenschatz, E.} 2009 Lagrangian properties of particles in turbulence. {\em Annu. Rev. Fluid Mech.\/} {\bf 41}, 375--404. \bibitem[Tsinober(2009)]{Tsinober2009book}{\sc Tsinober, A.} 2009 {\em An informal conceptual introduction to turbulence\/}. Springer Verlag. \bibitem[Xu {\em et~al.\/}(2006{\natexlab{{\em a\/}}})Xu, Bourgoin, Ouellette \& Bodenschatz]{Xu2006PRL}{\sc Xu, H.T., Bourgoin, M., Ouellette, N.T. \& Bodenschatz, E.} 2006{\natexlab{{\em a\/}}} High order lagrangian velocity statistics in turbulence. {\em Phys. Rev. Lett.\/} {\bf 96}~(2), 024503. \bibitem[Xu {\em et~al.\/}(2006{\natexlab{{\em b\/}}})Xu, Ouellette \& Bodenschatz]{Xu2006PRLb}{\sc Xu, H.T., Ouellette, N.T. \& Bodenschatz, E.} 2006{\natexlab{{\em b\/}}} Multifractal dimension of lagrangian turbulence. {\em Phys. Rev. Lett.\/} {\bf 96}~(11), 114503. \bibitem[Yeung(2002)]{Yeung2002ARFM}{\sc Yeung, P.K.} 2002 Lagrangian investigations of turbulence. {\em Annu. Rev. Fluid Mech.\/} {\bf 34}~(1), 115--142. \bibitem[Yu \& Meneveau(2010)]{Yu2010PRL}{\sc Yu, H.D. \& Meneveau, C.} 2010 Lagrangian refined kolmogorov similarity hypothesis for gradient time evolution and correlation in turbulent flows. {\em Phys. Rev. Lett.\/} {\bf 104}~(8), 084502. \end{thebibliography} \end{document} }

11 — 1401.5437

\caption{Decay of the largest values of the mutual information ($I_{i,j}>0.005$) for all molecules and inter-atomic distances investigated in this work. The symbols {\color{black}{$\Box$}}, {\color{red}{$\Diamond$}}, {\color{blue}{$\ocircle$}}, {\color{magenta}{$\times$}} and {\color{green0}{$\vartriangle$}} stand for r=0.8, 1$\rm r_e$, 1.5$\rm r_e$, 1.65$\rm r_e$, 2$\rm r_e$, respectively.}

12 — 1401.6318

\caption{(Color online) Absolute error in the norm $\Delta=\left\vert(\vert\Psi(\mathbf{r},t)|-|\Psi(\mathbf{r},t_{0})|)\right|$ on a logarithmic scale for different lower and upper bounds of the truncation error. The parameters of the Gaussian model problem are the same as in Fig.\ref{fig1}. The inset is a blow-up of the region at the beginning of the time propagation.} \label{fig2} \end{figure} % ------------------ FIG ---------------- In Fig.\ref{fig1}, we show the evolution of the time step in Fatunla's propagation for our Gaussian model problem. The pulse envelope is also plotted in arbitrary units to illustrate the duration of the pulse. We see that the time step becomes increasingly large after the end of the pulse, reaching values of around $2$ at the end of the total propagation ($500$ a.u. of time). It is clear that the most demanding part of the propagation, and therefore the most time consuming one, is during the interaction of the pulse with the system. This observation is important when it is necessary to propagate the wavefunction up to large distances after the end of the pulse, as is the case when the TSC method is used. In the results for the time step shown in Fig.\ref{fig1}, the latter one is adapted according to the condition $10^{-14}<\lvert \mathbf{T}_{n+1} \rvert < 10^{-9}$, that is, if the truncation error is lower than the lower bound $10^{-14}$ then we increase the time step, and if it is higher than the upper bound $10^{-9}$ it is decreased. With this choice, the overall conservation of the norm is about $10^{-5}$, which is enough for the model problem case we are studying. For many physical problems, this level of accuracy in the norm is sufficient but, if a higher accuracy is needed, then we might expect that it is sufficient to shift the bounds of the truncation error. However, as shown in Fig.\ref{fig2}, such a conclusion is not correct. In Fig.\ref{fig2}, we consider three different constraints on the truncated error and calculate on a logarithmic scale, the absolute error on the norm denoted by $\Delta$ as a function of time. This error is defined as the absolute value of the difference between 1 and the norm at time t . In these three cases, the time propagation is started with the same time step namely $10^{-3}$ a.u. This time step always increases while the truncated error is smaller than the prescribed lower bound and decreases if the truncated error is above the upper bound. In all three cases, we observe a significant loss of accuracy in $\Delta$ at the very beginning of the time propagation. As described by Madro\~nero and Piraux \cite{Madronero}, this is due to initially very small values of the denominators in Eq. (20) which leads to inaccurate values of the stiffness matrix elements and of the truncated error. This problem is therefore intrinsically related to Fatunla's method and leads to difficulties in correctly controlling the time step. In fact , if we keep the time step constant from the beginning, we have a much better control of $\Delta$. We have also checked that this is true even in the field free case. On the other hand, in general, we see from the inset in Fig.\ref{fig2} that we maintain a higher accuracy when the constraint on the truncated error is more severe. In addition, we also observe several small jumps in $\Delta$ the magnitude of which are much smaller than the jump in $\Delta$ at the beginning of the time propagation. We attribute these jumps to an accumulation of roundoff errors. Indeed, we expect more roundoff errors in the case the constraint on the truncated error is the strongest since a smaller time step leads to a larger amount of calculations. Note that the jump observed in the red continuous line corresponds to a change of only one digit in the accuracy of the norm. The overall relative accuracy we obtain even for the most severe constraint we use on the truncation error is of the order of $10^{-5}$. To achieve a greater accuracy, it is necessary to use a fixed and very small time step. These results show clearly that the achievable accuracy for the adaptive time step approach in Fatunla's method has a lower bound for a given initial time step. As a result, the use of Fatunla's method rests on a compromise between the computer time required and the accuracy needed. In the following, we consider the interaction of helium with a strong laser pulse. In that case, the accuracy on the norm reduces to about 4 significant digits when Fatunla's method is used. This prevents us to calculate the probability of single ionization in various channels where the latter one is less than $10^{-4}$ a.u. for field intensities currently used in the experiments.\\ In conclusion Fatunla's method allows one to treat stiff problems while fully exploiting the advantages of explicit schemes, namely that it only involves matrix vector multiplications. However it has its own limitations. \subsection{Krylov subspace method} \label{arnoldi} In this section we consider a powerful method to propagate the TDSE solution, which provides accuracy of solutions and stability of propagation. It uses projection techniques on Krylov subspaces \cite{Saad}. This approach was proposed by Arnoldi \cite{Arnoldi} in the calculation of the eigenstates of a matrix. Here we briefly recall the method used by Arnoldi as a time propagator \cite{Park}, to solve the differential equation (\ref{eq5}). Since the overlap matrix is positive-definite, we can use the Cholesky decomposition $\mathbf{B}=\mathbf{U}^{\dagger}\mathbf{U}$ to form an orthonormal basis defining the new coefficients $\mathbf{\Phi}=\mathbf{U}\boldsymbol{\Psi}$. The TDSE for these coefficients is written in the form, \begin{equation} \frac{\mathrm{d}\mathbf{\Phi}(t)}{\mathrm{d}t}=-\mathrm{i} \mathbf{\widehat{H}}(t)\mathbf{\Phi}(t) ,\label{eq22} \end{equation} where $\mathbf{\widehat{H}}=(\mathbf{U}^{\dagger})^{-1}\mathbf{H} \mathbf{U}^{-1}$. If we assume that the time interval is sufficiently small that the Hamiltonian may be treated as constant in time over a time step $\delta t$, it is trivial to demonstrate that Eq.(\ref{eq22}) has a solution given by \begin{equation} \mathbf{\Phi}(t+\delta t)=e^{-\mathrm{i} \mathbf{\widehat{H}}(t)\delta t} \mathbf{\Phi}(t). \label{eq23} \end{equation} If $\mathbf{\widehat{H}}$ is diagonalizable and can be written as $\mathbf{\widehat{H}}=\mathbf{E}\mathbf{\Lambda}\mathbf{E}^{-1}$, where $\mathbf{\Lambda}$ is a diagonal matrix with the eigenvalues $\lambda_{i}$ of $\mathbf{\widehat{H}}$ on the main diagonal and $\mathbf{E}$ is the matrix with the corresponding eigenvectors of $\mathbf{\widehat{H}}$ as its columns, then Eq.(\ref{eq23}) can be reexpressed as follows \begin{equation} \mathbf{\Phi}(t+\delta t)=\mathbf{E} e^{-\mathrm{i}\mathbf{\Lambda}(t)\delta t} \mathbf{E}^{-1} \mathbf{\Phi}(t) .\label{eq24} \end{equation} However, for very large $N$ this may be unnecessary and computationally very demanding. Instead, we can define the exponential in Eq.(\ref{eq23}) using a Taylor expansion of the form, \begin{equation} \mathbf{\Phi}(t+\delta t)=\left(\mathbf{I}-\mathrm{i} \delta t \mathbf{\widehat{H}}(t)+... + \frac{(-\mathrm{i} \delta t)^{k}}{k!} \mathbf{\widehat{H}}^{k}(t)+...\right)\mathbf{\Phi}(t). \label{eq25} \end{equation} We use the successive matrix products as a basis set forming a Krylov subspace spanned by $(m+1)$ linearly independent vectors, denoted by \begin{equation} K_{m+1}=span\{\mathbf{\Phi},\mathbf{\widehat{H}}\mathbf{\Phi},...,\mathbf{\widehat{H}}^{m}\mathbf{\Phi}\}. \label{eq26} \end{equation} To build the Krylov subspace, we first use Gram-Schmidt orthogonalization of the initial vectors $\{\mathbf{\Phi},\mathbf{\widehat{H}}\mathbf{\Phi},...,\mathbf{\widehat{H}}^{m}\mathbf{\Phi}\},$ to obtain an orthonormal basis $\{\mathbf{q}_{0},\mathbf{q}_{1},...,\mathbf{q}_{m}\}$. The procedure starts with $\mathbf{q}_{0}=\mathbf{\Phi}/\vert \mathbf{\Phi}\vert$, where the norm is defined as $\vert \mathbf{\Phi}\vert= \sqrt{\mathbf{\Phi}^{\dagger}\cdot\mathbf{\Phi}}$. The $\mathbf{q}_{k}$ are obtained by calculating $\mathbf{\widehat{H}}\mathbf{q}_{k-1}$ and then orthonormalizing each vector with respect to $\mathbf{q}_{0},...,\mathbf{q}_{k-1}$. If we define $\mathbf{Q}$ to be a matrix formed by the $m+1$ column vectors $(\mathbf{q}_{0},...,\mathbf{q}_{m})$, we finally get \begin{equation} \mathbf{\widehat{H}} \mathbf{Q}=\mathbf{Q} \mathbf{h},\label{eq27} \end{equation} giving \begin{equation} \mathbf{h}=\mathbf{Q}^{\dagger}\mathbf{\widehat{H}}\mathbf{Q}. \label{eq28} \end{equation} We see here that $\mathbf{h}$ is the Krylov subspace representation of the full Hamiltonian $\mathbf{\widehat{H}}$, and that in this procedure, we obtain simultaneoulsy the Krylov vectors $\mathbf{q}_{0},...,\mathbf{q}_{m}$. Arnoldi's algorithm is general and applies to non-hermitian matrices. It reduces the dense matrix $\mathbf{h}$ to an upper Hessenberg form, and in the particular case of hermitian matrices, to a symmetric tridiagonal form. In this latter case, Lanczos has shown that this matrix can be obtained by means of a simple recursion formula. However, this formula is known to be problematic when the size of the Krylov subspace is large because the orthogonality of the Krylov vectors is rapidly lost \cite{Saad}. It is the reason why we do not use this algorithm in the present case. Once we obtain the orthonormal Krylov subspace $\mathbf{Q}$ and the representation $\mathbf{h}$ of the Hamiltonian, it can be easily shown that Eq.(\ref{eq23}) can be written as \begin{equation} \mathbf{\Phi}(t+\delta t)=\mathbf{Q} e^{-\mathrm{i}\mathbf{h}\delta t} \mathbf{Q}^{\dagger} \mathbf{\Phi}(t). \label{eq29} \end{equation} The matrix $\mathbf{h}$ for all our case studies is tridiagonal, and its size is never bigger than $100 \times 100$, so the calculation of the exponential through direct diagonalization, as in Eq.(\ref{eq24}), is straightforward.\\ In actual numerical calculations, the Arnoldi algorithm \cite{Saad} requires some modifications. After a first calculation of a new Krylov vector $\mathbf{q}_{j+1}$, we ensure that the norm is equal to one, by re-checking the orthogonality against the previously calculated vectors, and perform again the Gram-Schmidt procedure if necessary. In principle the orthogonality condition determines the maximum size of the Krylov subspace and the algorithm can be used with a number $m-1$ of vectors. Also, if we start generating the Krylov vectors from the ground state of the system, then $\mathbf{\Phi}(t=t_{0})$ is an eigenstate of the Hamiltonian, making it impossible to build a linearly independent set of Krylov vectors. To solve this problem, instead of using the vector $\mathbf{\Phi}(t=t_{0})$ as a starting point, we use a modified vector $\mathbf{\Phi}(t=t_{0})+\boldsymbol{\varrho}$, with $\boldsymbol{\varrho}$ a vector of random entries no larger than $10^{-10}$.\\ % -------------------- FIG------------------ \begin{figure}[ttp] \begin{center} \includegraphics[width=11cm,height=9cm]{fig3.pdf} \end{center} \caption{Number of Krylov vectors required to obtain convergence of the final vector propagated for different values of the (fixed) time step. The parameters of the Gaussian model problem are the same as in Fig.\ref{fig1}.} \label{fig3} \end{figure} % ------------------ FIG ---------------- By construction (see Eq.\ref{eq23}), the stability function associated to the numerical time propagator based on Arnoldi's algorithm is given by $R(\lambda\delta t)=\exp(\lambda\delta t)$. As a result, it has exactly the same stability properties as Fatunla's algorithm. However, it is worth remembering that Eq.(\ref{eq23}) is only valid if the Hamiltonian is time independent. It is therefore a good approximation only for small values of $\delta t$. In the present case, there are two types of errors. The first one is directly related to Arnoldi's algorithm for the calculation of the exponential of a matrix. This type of error has been discussed in detail by Saad \cite{Saad2} and later on by Hochbruck and Lubich \cite{Hochbruck}. We have checked that this type of error is always negligible and does not depend on the time step. The second type of error is due to assuming that the Hamiltonian does not depend explicitly on time over the time step $\delta t$. We estimated this type of error by calculating $\|\mathrm{d}\mathbf{\Psi}/\mathrm{d}t+\mathrm{i}\mathbf{H}\mathbf{\Psi}\|$ and checked that, as expected, it is of the order $\delta t^2$. Another way to estimate this type of error is to compare our results with those obtained with an Arnoldi based method that takes explicitly into account the time dependence of the Hamiltonian. This can be done by using a Magnus expansion of the time evolution operator \cite{Magnus,Iserles}. However, this method requires very time consuming calculations beyond the scope of this contribution. On the other hand, as it was already noted by other authors \cite{Park}, enlarging the size of the Krylov space allows for larger time steps to be considered. In Fig.\ref{fig3}, we give the number of Krylov vectors necessary to obtain convergence of our results as a function of the time step used in the calculations for the case of the one-dimensional Gaussian model potential with the same parameters as in Fig.\ref{fig1}. In our calculations, the time step $\delta t$ is kept constant during the propagation. The choice of the optimal value of the time step and of the corresponding dimension of the Krylov space is therefore the result of a compromise while trying to reduce the computer time.\\ The innovative use of the Arnoldi method as an explicit approach offers then the convenience that we only require matrix-vector and scalar products, which then transforms the method in a time-efficient approach as is the case for Fatunla's method. Furthermore, this particular scheme is norm-conserving with the advantage of providing a check for the method, even though it also means that it is not easy to quantify the error in the calculation of the norm. In the following sections the accuracy of both methods is tested in various situations by using a high order predictor-corrector method. \subsection{A predictor-corrector method.} \label{predcor} In this subsection, we briefly describe the predictor-corrector (P-C) scheme we use to test the accuracy of both explicit methods described above. The predictor is either Fatunla's or Arnoldi's algorithm. The corrector is a fully implicit method of Runge-Kutta type which, here, is a four-stage Radau IIA method of order 7. \\ In a general Runge-Kutta method, the numerical solution $\mathbf{\Psi}_{n+1}$ of Eq.(\ref{eq8}) at a given time $t=t_{n+1}$ is obtained from the solution $\mathbf{\Psi}_n$ at time $t=t_n$ as \begin{equation} \boldsymbol{\Psi}_{n+1}=\boldsymbol{\Psi}_{n}+\delta t \sum_{i=1}^{s}b_{i}f(t_{i},\mathbf{Y}_{i}), \label{eq30} \end{equation} where $\delta t$ is the time step and $f(t_{i},\mathbf{Y}_{i})=-\mathrm{i} \mathbf{B}^{-1}\mathbf{H}(t_{i})\mathbf{Y}_{i}$ with $t_{i}=t_{n}+c_{i}\delta t$. $b_{i}$ and $c_{i}$ are coefficients defining the Runge-Kutta method for a number $s$ of stages. We assume that the solution vector $\mathbf{\Psi}$ is of dimension $N$. The quantities $\mathbf{Y}_{i}$ estimate the solution $\mathbf{\Psi}$ at the intermediate time $t_i$. They are obtained by solving the following linear ($sN\times sN$) system of equations \begin{equation} \mathbf{Y}_{i}=\boldsymbol{\Psi}_{n}+\delta t \sum_{k=1}^{s}a_{ik}f(t_{k},\mathbf{Y}_{k}), \label{eq31} \end{equation} where the $a_{ik}$ are again given by the method. Solving such system represents the main difficulty of an implicit Runge-Kutta scheme. If this scheme is used for the corrector, we could in principle avoid solving such system of equations by using an iterative procedure in which we replace the vector $\mathbf{Y}_k$ in the right hand side of Eq.(\ref{eq31}) by $\mathbf{Y}^{(j-1)}_k$ where $j$ gives the order of the iteration process. At the order $0$, $\mathbf{Y}^{(0)}_k$ is provided by the predictor. However, such an iterative procedure is not stable. Instead, we follow a different iterative procedure that has been developed by van der Houwen and Sommeijer \cite{Houwen1,Houwen2}. By introducing a diagonal matrix whose entries are calculated to guarantee optimum stability properties, they transform the ($sN\times sN$) system (\ref{eq31}) into a set of uncoupled ($N\times N$) systems of equations that can be solved in parallel at each iteration. More precisely, they rewrite Eq.(\ref{eq31}) as follows \begin{equation} \mathbf{Y}_{i}^{(j)}-\delta t \, d_{ii} \, f(t_{i},\mathbf{Y}_{i}^{(j)})=\boldsymbol{\Psi}_{n}+ \delta t \sum_{k=1}^{s}(a_{ik}-d_{ik})f(t_{k},\mathbf{Y}_{k}^{(j-1)}), \label{eq32} \end{equation} where the $d_{ik}$ are the entries of the diagonal matrix. The iterations in Eq.(\ref{eq32}) start with $\mathbf{Y}_{i}^{(0)}$, provided by the predictor. The iteration scheme is performed until a value $j=$\textbf{max}$_{cor}$ for which we have convergence. Then we can replace $\mathbf{Y}_{i}=\mathbf{Y}_{i}^{(m)}$ in Eq. (\ref{eq30}) to obtain the solution at $t=t_{n+1}$. Once we have calculated $\mathbf{\Psi}_{n+1}$, we can evaluate its norm and use its conservation as a criterion to monitor the size of the time step.\\ % -------------------- FIG------------------ \begin{figure}[h] \begin{center} \includegraphics[width=11cm,height=8cm]{fig4.pdf} \end{center} \caption{(Color online) Number of iterations in the Bi-CGSTAB and their multiplicities during the propagation with the electromagnetic pulse. The parameters of the Gaussian model problem are $V_{0}=4$ a.u., $\beta=0.1$ a.u., with a pulse of peak intensity $I=10^{16}$ Watt/cm$^2$, photon energy $\omega=0.5$ a.u. and duration of 8 optical cyles.} \label{fig4} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[b!] \begin{center} \includegraphics[width=11cm,height=8cm]{fig5.pdf} \end{center} \caption{(Color online) Time step evolution for P-C method of time propagation during the propagation with the electromagnetic pulse. The parameters of the Gaussian model problem are the same as in Fig.\ref{fig4}.} \label{fig5} \end{figure} % ------------------ FIG ---------------- Using the P-C method requires solving a large number of $(N\times N)$ systems of equations. To solve these systems, we use an iterative method known as the biconjugate gradient stabilized method (Bi-CGSTAB) \cite{Vorst}. In order to reduce drastically the number of iterations, we use a pre-conditioner based on an incomplete LU factorization. In Fig.\ref{fig4}, we consider the case of our one-dimensional Gaussian model potential with $V_0=4$ a.u. and $\beta=0.1$ a.u. and a pulse of frequency $\omega=0.5$ a.u., duration of 8 optical cycles and peak intensity $ I=10^{16}$ Watt/cm$^2$. We show, in this Fig.\ref{fig4}, the multiplicity as a function of the number of iterations in the Bi-CGSTAB during the interaction. By multiplicity, we mean the number of times a given number of iterations is repeated during the whole propagation. It can be seen that without pre-conditioner, the number of iterations can grow significantly before reaching convergence, while using the preconditioned Bi-CGSTAB, the number of iterations is maintained below five. This reduces the computational time needed by 25\%. However, care must be taken when including a pre-conditioner, since, by increasing the number of operations, it may increase the computational times even though it accelerates convergence. As mentioned above, the corrector scheme is iterated up to $j=$\textbf{max}$_{cor}$ where convergence is achieved. In Fig.\ref{fig5} we plot the time evolution of the time step during the propagation for different values of the maximum of iterations \textbf{max}$_{cor}$ in the corrector . We see here that, as we increase this maximum number of iterations, the value of the time step becomes larger. The relative error in the norm which is the same for all the calculations is of the order of $10^{-11}$. Moreover, the computational time with \textbf{max}$_{cor}$=100 is half the time consumed for \textbf{max}$_{cor}$=10 because it allows to use a much larger time step. It is therefore advisable to use large values of \textbf{max}$_{cor}$ to speed up the calculations. \section{Results} \subsection{Model potential} In this section we first present results for our case study of the one-dimensional Gaussian potential taking $V_0=1$ and $\beta=1$. The electron wavepacket is developed in a basis of 200 B-splines and we use the time scaled coordinate method during the propagation \cite{Hamido}. % -------------------- FIG------------------ \begin{figure}[!h] \begin{center} \includegraphics[width=10cm,height=6.6cm]{fig6.pdf} \end{center} \caption{(Color online) Energy distribution for the Gaussian model potential. The time propagation uses Fatunla's propagator with adaptive time step and Arnoldi's propagator with five Krylov vectors and a fixed time step $\delta t = 0.3$ a.u. The parameters of the model problem are $V_0=1$ and $\beta=1$ with a pulse of peak intensity I=$10^{14}$ Watt/cm$^2$, photon energy $\omega=0.7$ a.u. and a duration of $10$ optical cycles. The relative difference between both curves is of the order of $10^{-3}$.} \label{fig6} \end{figure} % -------------------- FIG------------------ We run our codes on an INTEL XEON 2.33 GHz Processor 51.40 (32 GB Ram). We choose a pulse of frequency $\omega=0.7$ a.u. and a full duration of 90 a.u. of time which corresponds to 10 optical cycles. The peak intensity I = $10^{14}$ Watt/cm$^2$. % -------------------- FIG------------------ \begin{figure}[!h] \begin{center} \includegraphics[width=9.5cm,height=6.8cm]{fig7.pdf} \end{center} \caption{(Color online) Energy distribution for the Gaussian model potential. The time propagation uses Fatunla's propagator with adaptive time step and Arnoldi's propagator with $20$ Krylov vectors and a fixed time step $\delta t = 0.1$ a.u. The parameters of the model problem are $V_0=1$ and $\beta=1$ with a pulse of peak intensity $I=10^{14}$ Watt/cm$^2$, photon energy $\omega=0.1$ a.u. and duration of $10$ optical cycles.The relative difference between both curves is of the order of $10^{-3}$.} \label{fig7} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[!b] \begin{center} \includegraphics[width=12cm,height=10cm]{fig8.pdf} \end{center} \caption{Energy distribution for the Gaussian model potential. The time propagation uses Fatunla's propagator with adaptive time step and the predictor-corrector method. The parameters of the model problem are $V_0=4.0$ and $\beta=0.1$ with a pulse of peak intensity $I=10^{16}$ Watt/cm$^2$, photon energy $\omega=0.5$ a.u. and duration of $8$ optical cycles.} \label{fig8} \end{figure} % ------------------ FIG ---------------- In Fig.\ref{fig6}, the energy distribution is calculated by propagating the scaled wavepacket to a stationary state until a time of 1500 a.u. when convergence is achieved. The results shown are obtained using the two explicit propagators. Both methods converge to the same result but Fatunla's propagator uses 2.3 s of computer time with an adaptive time step while Arnoldi's propagator using five Krylov vectors and a fixed time step $\delta t = 0.3$ a.u. takes 6.2 s. For these values of intensity and frequency both methods give easily the correct result. However Arnoldi's method performs poorly from a computer time point of view. This can be understood by referring to Fig.\ref{fig1} where we show that Fatunla's propagator allows the use of ever larger time steps, particularly during the propagation after the end of the pulse, while Arnoldi's propagator keeps the same time-step throughout the propagation.\\ % -------------------- FIG------------------ \begin{figure}[ttp] \begin{center} \includegraphics[width=12cm,height=10cm]{fig9.pdf} \end{center} \caption{Energy distribution for the Gaussian model potential. The time propagation uses Arnoldi's propagator with fixed time step and the P-C method with adaptive time step. The parameters of the Gaussian model problem are as in Fig.\ref{fig8}.} \label{fig9} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ To check how these methods behave in a more challenging case, we consider the same model potential with a pulse of frequency 0.1 a.u. with the same number of optical cycles and peak intensity. In this case, the total pulse duration is equal to 630 a.u. We see in Fig.\ref{fig7} that both methods give identical results. These results are obtained after propagating the wavepacket up to a time of 2500 a.u. The running time with Fatunla's propagator is equal to 958.27 s while in this case, Arnoldi's propagator performs better using 553.69 s for a subspace of 20 Krylov vectors and a fixed time step $\delta t = 0.1$ a.u. It can be seen that in general, Arnoldi's propagator performs better than Fatunla's propagator for long pulses.\\ To further probe these methods we increase the number of bound states supported by our potential by choosing ${V_0} = 4$ and $\beta=0.1$. The pulse has a frequency $\omega=0.5 $ a.u. with a duration of 100.53 a.u. that corresponds to 8 optical cycles. The peak intensity I = $10^{16}$ Watt/cm$^{ - 2}$. In this case, we use 1700 B-splines to propagate the wavepacket up to a time of 5000 a.u. Fig.\ref{fig8} shows the energy distribution obtained using Fatunla's propagator (straight line) and the predictor-corrector scheme (squares), which is used to test the accuracy of Fatunla's method. Comparison of these two methods shows that Fatunla keeps the accuracy in the results down to a value of $10^{-5}$ a.u. for the energy distribution. The TSC approach is used with an asymptotic scaling factor of 0.1. Fatunla's propagator takes 379.36 s while the P-C method with adaptive time step takes 1801.49 s. It is clear that Fatunla uses remarkably less computer time and works as long as the accuracy required is up to six digits. Fig.\ref{fig9} shows the energy distribution obtained with Arnoldi's propagator for the same parameters as in Fig.\ref{fig8}. We show the results obtained with Arnoldi's approach and two different fixed time steps and compare these results with those obtained with the P-C scheme. The Krylov subspace contains 20 vectors and the wavepacket is again propagated up to 5000 a.u. The circles show the results for a time step $\delta t=0.03$ a.u. and the stars for $\delta t=0.3$ a.u. We note that increasing the time step leads to less accurate results by comparison with the P-C method. Arnoldi scheme takes 5957.19 s for a time step of 0.03 a.u. and 598.11 s for a time step of 0.3 a.u.\\ \subsection{Hydrogen Atom} We now apply these methods to the more complex case of the interaction of hydrogen atom with a cosine square laser pulse. We use a spectral method based on the expansion of the wavefunction in a basis of Coulomb Sturmian functions \cite{Madronero}, without implementing the TSC method. Unless otherwise stated, we performed all calculations on a laptop (with an INTEL core 2 duo processor of 2.4 GHz). The first pulse we use has a frequency of 0.7 a.u., a duration of 10 optical cycles and an intensity I=10$^{14}$ Watt/cm$^2$, as in the case of Fig.\ref{fig6}. In these rather simple conditions, we use 10 angular momenta. The non-linear parameter $\kappa$ of the Coulomb Sturmian functions is taken equal to 0.3 a.u. Fatunla's and Arnoldi's algorithms produce the converged energy distribution as shown in Fig.\ref{fig10}. The calculations carried on with Fatunla's propagator and an adaptive time step take 10.50 s of computer time. The integration performed with Arnoldi's method takes 13.72 s. For a time step of $\delta t=0.05$ a.u. and 100 Coulomb Sturmian functions per angular momentum, it needs only 5 Krylov vectors. In this case Arnoldi's method is slower than Fatunla's method. This is related to the number of basis-set functions used. As this number increases, higher eigenvalues are generated in the Hamiltonian spectrum thereby increasing the stiff character of the system of equations to solve. In that case, more Krylov vectors have to be included to maintain the accuracy of the results. \begin{figure}[!ht] \begin{center} \includegraphics[width=10cm,height=7cm]{fig10.pdf} \end{center} \caption{(Color online) Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Fatunla's and Arnoldi's propagators are used. The pulse has a peak intensity I=10$^{14}$ Watt/cm$^2$, a frequency $\omega=0.7$ a.u. and a duration of 10 optical cycles. The basis-set of functions used is a set of 100 Coulomb Sturmian functions per angular momentum. Ten angular momenta are included and the non-linear parameter $\kappa$ of the Coulomb Sturmian functions is equal to 0.3. The Arnoldi propagator uses 5 Krylov vectors and a time step of $\delta t = 0.05$ a.u.The relative difference between both curves is of the order of $10^{-3}$.} \label{fig10} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[!ht] \begin{center} \includegraphics[width=10cm,height=6.8cm]{fig11.pdf} \end{center} \caption{(Color online) Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Arnoldi's propagator is used. The pulse parameters are as in Fig.\ref{fig10}, using 100 Coulomb Sturmian functions per angular momentum. Ten angular momenta are included and the non-linear parameter $\kappa$ of the Coulomb Sturmian functions is equal to 0.3. For a time step of $\delta t = 0.05$ a.u., we compare results when 5 and 4 Krylov vectors are used.} \label{fig11} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ %\begin{figure}[h] %\begin{center} %\includegraphics[width=10cm,height=7cm]{fig12.pdf} %\end{center} %\caption{(Color online) Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Arnoldi's propagator is used. The pulse parameters are as in Fig.\ref{fig10}. We use 100 Coulomb Sturmian functions per angular momentum. Ten angular momenta are included. For a time step of $\delta t = 0.05$ a.u., we compare the results obtained for two different values of the non-linear parameter of the Coulomb Sturmian functions, $\kappa=0.3$ and 0.4 while the number of Krylov vectors is fixed to 5.} %\label{fig12} %\end{figure} %% ------------------ FIG ---------------- %% -------------------- FIG------------------ %\begin{figure}[!h] %\begin{center} %\includegraphics[width=10cm,height=7cm]{fig13.pdf} %\end{center} %\caption{(Color online) Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Arnoldi's propagator is used. The pulse parameters are as in Fig.\ref{fig10}. We use 150 Coulomb Sturmian functions per angular momentum. Ten angular momenta are included and the non-linear parameter $\kappa=0.3$. For a time step of $\delta t = 0.05$ a.u., we compare results obtained with 5 Krylov vectors as in Fig.\ref{fig10} and 8 Krylov vectors, which is now the minimum number needed to reach convergence.} %\label{fig13} %\end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[h] \begin{center} \includegraphics[width=10cm,height=7.2cm]{fig14.pdf} \end{center} \caption{(Color online) Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Arnoldi's propagator is used. The pulse has a peak intensity I=10$^{14}$ Watt/cm$^2$, a frequency $\omega=0.114$ a.u. and a duration of 20 optical cycles. We use a set of 600 Coulomb Sturmian functions per angular momentum. Ten angular momenta are taken into account and the non-linear parameter of the Coulomb Sturmian functions $\kappa=0.3$. The Arnoldi propagator uses 25 Krylov vectors and a time step of $\delta t = 0.05$ a.u. The relative difference between both curves is of the order of $10^{-3}$.} \label{fig14} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[!ht] \begin{center} \includegraphics[width=10cm,height=7.2cm]{fig15.pdf} \end{center} \caption{(Color online)Energy distribution resulting from the interaction of the hydrogen atom with a cosine square pulse. Arnoldi's propagator is used. The pulse has a peak intensity I=10$^{14}$ Watt/cm$^2$, a frequency $\omega=0.0228$ a.u. and a duration of 4 optical cycles. The basis-set of functions used is a set of 1200 Coulomb Sturmian functions per angular momentum. 80 angular momenta are taken into account and the non-linear parameter $\kappa$ of the Coulomb Sturmian functions is equal to 0.3. The Arnoldi propagator uses 70 Krylov vectors and a time step of $\delta t = 0.05$ a.u.} \label{fig15} \end{figure} % ------------------ FIG ---------------- In Fig.\ref{fig11} we illustrate the effect of reducing the number of Krylov vectors $n_k$ from 5 to 4. It is surprising to see that the propagator gives a completely flat spectrum when the dimension of the Krylov space is insufficient. Fig.\ref{fig11} shows that for a basis set of 100 Coulomb Sturmian functions per angular momentum, accurate results for the energy distribution require a minimum of 5 Krylov vectors.\\ These calculations performed in a Coulomb Sturmian basis can be further tested by varying their non-linear parameter $\kappa$. If instead of using $\kappa=0.3$, we use $\kappa=0.4$, all the other parameters remaining the same, we again obtain a completely flat energy distribution. By increasing the value of the non-linear parameter $\kappa$, the value of the eigenenergies increases thereby increasing the stiff character of the problem. To successfully reproduce an accurate energy distribution we now would need to increase the number of Krylov vectors. If on the other hand, we keep the value of $\kappa$ equal to 0.3 and increase the number of basis functions, converged results are only obtained when 8 Krylov vectors are used. The increase in the number of Coulomb Sturmians generates higher eigenenergies thereby increasing again the stiff character of the system. The eigenvalues of matrix \textbf{h} range from the eigenvalue of the initial state (by construction) to approximately the highest one of matrix \textbf{H}. In summary, any change which results in a higher maximum eigenvalue for \textbf{H} necessitates an increase in the number of Krylov subspace vectors required for convergence. It is interesting to note that to get an accurate spectrum, one of the eigenvalues of \textbf{h} must converge to 0.2 which corresponds to the position of the maximum of the spectrum which is what we expect from energy conservation ($0.2=-0.5+\omega$). If none of the eigenvalues converges to 0.2, the spectrum is completely flat because all the eigenvalues of \textbf{h} which are usually very high except the first one, do not contribute significantly to the spectrum. In addition, it is important to stress that decreasing the time step does not modify the minimum number of Krylov vectors to be used.\\ In Fig.\ref{fig14} we compare the performance of Arnoldi's and Fatunla's methods for a more difficult case. We consider a pulse of frequency $\omega=0.114$ a.u. and a duration of 20 optical cycles. The pulse intensity is the same as before, I=10$^{14}$ Watt/cm$^2$. We use a basis set of 600 Coulomb Sturmian functions per angular momentum. 10 angular momenta are included in the calculations and the non-linear parameter $\kappa=0.3$. Both energy distributions agree but Fatunla's scheme, which needs a very small time step, takes 66004 s of computer time while Arnoldi's method takes 1419 s with 25 Krylov vectors and a time step of 0.05 a.u. This case illustrates clearly that Arnoldi's algorithm copes in an efficient way with the stiffness of the problem by increasing the size of the Krylov subspace.\\ In Fig.\ref{fig15} we show results obtained for the challenging case of a pulse of very low frequency $\omega=0.0228$ a.u. and a duration of 4 optical cycles for the same intensity as before. To reproduce the energy distribution we need to use 1200 Coulomb Sturmian functions per angular momentum. 80 angular momenta are included in the calculations and the non-linear parameter $\kappa$ of the Coulomb Sturmian functions is equal to 0.3. For this rather stiff problem Arnoldi's algorithm has to include a minimum of 70 Krylov vectors for a time step $\delta t = 0.05$ a.u. The calculation takes 24 hours on an 8 processor cluster using OpenMP. Fatunla's algorithm also reproduces the same energy distribution but the computer time used is more than four times larger. In fact, we observe that for larger scale problems where the degree of stiffness is important, Fatunla's method requires time steps that become prohibitively small thereby increasing the computational time. \subsection{Helium Atom} % -------------------- FIG------------------ \begin{figure}[!ht] \begin{center} \includegraphics[width=12cm,height=8cm]{fig16.pdf} \end{center} \caption{Single ionisation spectrum resulting from the interaction of an helium atom with a short cosine square pulse. Arnoldi's propagator is used. The pulse has a peak intensity of I=10$^{14}$ Watt/cm$^2$, a frequency $\omega=2.1$ a.u. and a duration of 6 optical cycles. The basis-set of functions used is a set of 140 B-spline functions of order 7 per electron angular momentum. The total angular momentum L=0,1,2 and the maximum value of the individual electron angular momentum is three. The box size is 150 a.u. The Arnoldi propagator uses 40 Krylov vectors.} \label{fig16} \end{figure} % ------------------ FIG ---------------- In this subsection, we show briefly results for the single ionization of helium by an intense electromagnetic pulse as an example of a more challenging problem. Following the remarks above, we perform here the calculations using only Arnoldi's algorithm. As mentioned before, Fatunla's algorithm is not accurate enough to calculate cross sections in various single ionization channels. The pulse has a peak intensity I=10$^{14}$ Watt/cm$^2$, a frequency $\omega=2.1$ a.u. and a duration of 6 optical cycles. The wavefunction is expanded in a basis set that uses 140 B-spline functions of order 7 per electron angular momentum \cite{Bachau}. Three values of the total angular momentum (L=0,1,2) are taken into account and the maximum value of the individual electron angular momentum is three. The box size is 150 a.u. The step size during the interaction with the pulse is fixed at 0.01 a.u., while after the interaction the propagation used a step size of 1 a.u. The calculations are performed with 40 Krylov vectors. It takes 31 hours to run on a cluster with 10 Intel Xeon L5520 2.26 GHz processors using MPI (Message Passing Interface) and 3 GB of RAM per processor. Fig.\ref{fig16} shows the results obtained for the energy distribution of the single ionization of helium. As expected we observe a dominant peak at 1.2 a.u. which corresponds to the energy conservation. The spectrum is obtained by projecting the wave packet after the end of the pulse on a product of a Coulomb wave of the screened nucleus times a bound state of He$^{+}$. \subsection{Quantum dot} In this last section, we consider a different problem where the choice of a very efficient explicit time propagator turns out to be crucial. The system under consideration is a model for a planar two-electron quantum dot with an anharmonic confining potential. The properties of quantum dots have great resemblance to those of atoms or molecules. Optical lattices, which can be viewed as an array of quantum dots, and well-approved methods from semiconductor physics make quantum dots easily accessible. A confinement of the electrons to a two-dimensional plane is justified, in particular for solid state quantum dots, where the electron gas is localized on a parallel plane between two layers of different semiconductors. The Hamiltonian for this problem is of the form \begin{equation} H_{\varepsilon} = {H_1} + {H_2} + {V_{{\mathop{\rm int}} }},\label{eq33} \end{equation} where the indices $1$ and $2$ refer to the two electrons. ${V_{{\mathop{\rm int}} }} =\displaystyle{ \frac{1}{{\,{r_{1\,2}}}}}$, with $ {{r_{1\,2}}}$ being the inter-electronic distance. The Hamitonians $H_j$ are given by, \begin{equation} {H_j} = \frac{1}{2} \mathbf{p}_j^{{\kern 1pt} 2} + \frac{{{\omega ^2}}}{2} \mathbf{r}_j^{{\kern 1pt} 2} + \varepsilon {\left( { \mathbf{r}_j^{{\kern 1pt} 2}} \right)^2}, \label{eq34} \end{equation} with $\omega$ the harmonic frequency and $\varepsilon$ the strength of the anharmonic perturbation. $\mathbf{r}_j$ and $\mathbf{p}_j$ are the coordinate and momentum of electron $j$, respectively. For $\varepsilon\equiv 0$ our model coincides with the well-known Hooke's atom, which is separable in the centre-of-mass and relative coordinates. The Schr\"odinger equation can be regularized \cite{Schroeter1} using the Jacobian of a suitable parabolic coordinate transformation. We then write the resulting equation in terms of circular creation and annihilation operators. A set of selection rules is obtained determining the coupling between basis states and the matrix elements, according to the principal quantum numbers of the harmonic oscillators. The TDSE, \begin{equation} H\;\Psi \left( {{{\bf{r}}_{\,1}},{{\bf{r}}_{\,2}},t} \right) = \mathrm{i} \frac{\partial }{{\partial \,t}}\;\Psi \left( {{{\bf{r}}_{\,1}},{{\bf{r}}_{\,2}},t} \right), \label{eq35} \end{equation} is solved to obtain $\Psi \left({{{\bf{r}}_{\,1}},{{\bf{r}}_{\,2}},t} \right)$, with $H$ given in Eq.(\ref{eq33}). The question of decoherence of these quantum states can be studied through the quantum fidelity, which gives the overlap of the solutions of the TDSE, with and without the potential ${V_{{\mathop{\rm anharmonic}} } =\displaystyle{ \varepsilon {\left(\left( { \mathbf{r}_1^{{\kern 1pt} 2}} \right)^2+\left( { \mathbf{r}_2^{{\kern 1pt} 2}} \right)^2\right)}}}$. The perturbation potential ${V_p =\displaystyle{ \left( { \mathbf{r}_1^{{\kern 1pt} 2}} \right)^2+\left( { \mathbf{r}_2^{{\kern 1pt} 2}} \right)^2}}$ breaks the separability of Hooke's atom. We note that the Hamiltonian $H_{\varepsilon}$ in this case is not explicitly dependent on time and so it is different in nature to the Hamiltonians that we treated in previous examples. For a general Hamiltonian $H_0$ and a small real parameter $\varepsilon$ that represents the strength of the perturbation, we write \begin{equation}{H_\varepsilon } = {H_0} + \varepsilon{\kern 1pt}{V_{{\mathop{\rm p}.} }}\label{eq36} \end{equation} The quantum fidelity $F_\varepsilon$ at time $t$ is defined as, \begin{equation}{F_\varepsilon }\left(t \right)= {\left|{\left\langle{{\Psi _0}\left(t \right)} \right|\left. {{\Psi _\varepsilon }\left(t \right)} \right\rangle} \right|^2},\label{eq37} \end{equation} where $\Psi _\varepsilon$ and $\Psi _0$ are the quantum states propagated with Eq.(\ref{eq35}) for a perturbed and non-perturbed Hamiltonian, respectively. We can expand the quantum fidelity in terms of the perturbation parameter $\varepsilon$ \cite{Gorin}, as, \begin{equation}{F_\varepsilon }\left(t \right)= 1 - \chi \left(t \right){\varepsilon ^2} + O\left({{\varepsilon ^4}} \right),\label{eq38} \end{equation} with $ \chi \left( t \right)$ being the quantum susceptibility. Taking the two first terms, we evaluate ${F_\varepsilon }\left(t \right)$ up to order ${\varepsilon ^2}$, valid near unity. % -------------------- FIG------------------ \begin{figure}[!b] \begin{center} \includegraphics[width=12cm,height=8cm]{fig17.pdf} \end{center} \caption{(Color online) Susceptibility $ \chi \left( t \right)$ calculated for a quantum dot with $\omega=$1.0 a.u., using the Arnoldi's propagator. We took 5 Krylov vectors and compared results for two different values of the time step.} \label{fig17} \end{figure} % ------------------ FIG ---------------- % -------------------- FIG------------------ \begin{figure}[!ht] \begin{center} \includegraphics[width=12cm,height=8cm]{fig18.pdf} \end{center} \caption{(Color online) Susceptibility $ \chi \left(t \right)$ calculated for a quantum dot with $\omega=$1.0 a.u., using the Arnoldi's propagator. We take two different combinations of time step and of the dimension of the Krylov subspace used, to illustrate the compromise between these two quantities.} \label{fig18} \end{figure} % ------------------ FIG ---------------- For our particular case $H_0=H_{\varepsilon=0}$ and consequently $V_{{\mathop{\rm p}} }=\displaystyle{ {\left(\left( { \mathbf{r}_1^{{\kern 1pt} 2}} \right)^2+\left( { \mathbf{r}_2^{{\kern 1pt} 2}} \right)^2\right)}}$. The observable calculated in this problem is the susceptibility $\chi \left( t \right)$ and we take the harmonic frequency to be $\omega=$1.0 a.u. and the perturbation parameter to be $\varepsilon=10^{-5}$. We study the evolution of the initial bound state of energy E=7 a.u. and vanishing angular momentum, singlet state with even parity \cite{Schroeter2}. The total number of functions in the basis set is 2370. The integration of the TDSE was first attempted using Fatunla's method. The stiffness of this problem forces the adaptive time step to become excessively small (of the order of $10^{-5}$) so that the computer time needed by the method becomes of the order of several days instead of seconds. Furthermore the accuracy necessary to represent the effect of very small perturbations on the system could not be achieved. As a consequence we used Arnoldi's integrator, testing different combinations of the values of the time step and of the dimension of the Krylov subspace. It is worth stressing that the time evolution operator calculated within Arnoldi's method is essentially exact since the total Hamiltonian is time independent. However, the stiffness of the problem which is very strong because of the anharmonic character of the potential is expected to impose important constraints on the time step. In Fig.\ref{fig17} we show results for the quantum susceptibility using 5 Krylov vectors and two different time steps. In order to get converged results, this shows that we need a time step of at least $\delta t = 10^{-4}$ a.u., leading to a computational time of 4 hours. The same calculation performed with the P-C method took 17 days, 8 hours and 29 minutes. Fig.\ref{fig18} shows results for the observable $\chi \left(t \right)$ under the same conditions as in Fig.\ref{fig17} but using Krylov subspaces of higher dimension ($n_k=7$ and $n_k=9$). For $n_k=7$ converged results were obtained with a step size of $\delta t = 5 \times10^{-4}$ a.u. leading to a computational time of 45 minutes. However this figure illustrates the compromise to be achieved between the size of the time step used and the dimension of the Krylov subspace. For $n_k=9$, a time step of $\delta t = 10^{-3}$ a.u. leads to a calculation taking 30 minutes of computer time only. The choice of the optimal value of time step and of the Krylov subspace dimension needs to be balanced. This means to search for the optimal larger value of the time step for which the propagation will take less iterations. These calculations performed with the P-C method take 9 days, 17 hours and 14 minutes for $n_k=7$ and $\delta t = 5 \times10^{-4}$ and 8 days, 5 hours and 43 minutes for a.u. $n_k=9$ and $\delta t = 10^{-3}$ a.u. The computer used in these calculations was a single core of a Intel(R) Core (TM) 2 Quad CPU Q 9400(2.66 GHz) with 8 GB main memory. \section{CONCLUSIONS} In this contribution, we addressed the problem of the numerical integration of the time-dependent Schr\"odinger equation describing physical processes whose complexity requires the use of state of the art methods. The problem can be reduced to the solution of a system of first order differential equations. The main difficulties we have to face are the size of the system and its stiff character which results from the presence of very high energy eigenvalues in the Hamiltonian spectrum. These difficulties impose important constraints on the choice of the time propagator. Given the size of the system, this time propagator must be explicit. This means that it involves only matrix-vector products instead of solving large system of algebraic equations at each time step as is the case for implicit methods. In addition, this propagator must have optimum stability and accuracy properties to cope with the stiffness of the system. We have analyzed and compared the performance of two one-step explicit time propagators namely Fatunla's and Arnoldi's algorithms. It turns out that both of these methods share the same optimum stability properties. Nevertheless, we show that their accuracy properties differ significantly in most of the problems that we treat here. As a matter of fact, the accuracy of the method depends essentially on the stiffness of the system to solve which determines the appropriate choice of the propagator.\\ In all the problems considered here, the relative accuracy of Fatunla's method is always limited to about $10^{-6}$. In some cases, this might be sufficient but we should not forget that when the degree of stiffness increases, the adaptive time step becomes excessively small making the method inapplicable. By contrast, highly accurate results are obtained with Arnoldi's algorithm in all cases treated here. However, for a given time step, there is a minimal number of Krylov vectors to take into account. If the actual number used is smaller than this minimal number, generally there is an abrupt transition and the results are wrong giving a flat spectrum (in some cases this transition is not so abrupt but is rapid nevertheless.) On the other hand, when the degree of stiffness is high, this minimal number may become very large thereby imposing strong limitations on the applicability of the method. This is the case when the spacing between grid points becomes very small or, for spectral methods, when the size of the basis set is very large. In applying Arnoldi's scheme, it is therefore important to try to reduce the stiffness as much as possible. An obvious way to achieve this is to move to the atomic basis in which the Hamiltonian is diagonal and to eliminate the highest energy eigenvalues which, in principle do not play any physical role. In that case however, the ac-Stark shift of the levels will not be evaluated accurately. In addition, our calculations in the case of the Gaussian potential model clearly show that the energy electron spectrum calculated with Arnoldi's algorithm deteriorates. \section{ACKNOWLEDGEMENTS} A.L.F. gratefully acknowledges the financial support of the IISN (Institut Interuniversitaire des Sciences Nucl\'eaires) through contract No. 4.4.504.10, "Atoms, ions and radiation; Experimental and theoretical study of fundamental mechanisms governing laser-atom interactions and of radiative and collisional processes of astrophysical and thermonuclear relevance". F.M.F. and P.F.O'M thank the Universit\'e catholique de Louvain (UCL) for financially supporting several stays at the Institute of Condensed Matter and Nanosciences of the UCL. They also thank The european network COST (Cooperation in Science and Technology) through the Action CM1204 "XUV/X-ray light and fast ions for ultrafast chemistry (XLIC) for financing one short term scientific mission at UCL. Computational resources have been provided by the supercomputing facilities of the UCL and the Consortium des Equipements de Calcul Intensif en F\'ed\'eration Wallonie Bruxelles (CECI) funded by the Fonds de la Recherche Scientifique de Belgique (F.R.S.-FNRS) under convention 2.5020.11.\newpage \begin{thebibliography}{99} \bibitem{Muller} H.G. Muller, P. Agostini and G. Petite, in \emph{Atoms in Intenses Laser Fields}, edited by M. Gavrila (Academic Press, New York, 1992). \bibitem{L'Huillier} A. L'Huillier, L. Lompr\'e, G. Mainfray and C. Manus, in\emph{Atoms in Intenses Laser Fields}, edited by M. Gavrila (Academic Press, New York, 1992). \bibitem{Hansch} P. Hansch, M.A. Walker and L.D. Van Woerkom, Phys. Rev. A \textbf{54}, R2559 (1996). \bibitem{Antoine} Ph. Antoine, D.B. Milosevic, A. L'Huillier, M. Gaarde, P. Sali\`eres and M. Lewenstein, Phys. Rev. A\textbf{56}, 4960 (1997). \bibitem{Corkum} P.B. Corkum, Phys. Today. 64, No. 336 (2011). \bibitem{Borot} A. Borot, A. Malvache, X. Chen, A. Jullien, J.-P. Geindre, P. Audebert, G. Mourou, F. Qu\'er\'e and R. Lopez-Martens, Nature Physics\textbf{8}, 416 (2012). \bibitem{Burke} P.G. Burke, in \emph{Atomic, Molecular and Optical Physics Handbook} edited by G.W.F. Drake (AIP Press, New York, 1996). \bibitem{Joachain} C.J. Joachain, \emph{Quantum Collision Theory} (North-Holland Publishing Company Amsterdam, Oxford 1975). \bibitem{Sidky} E.Y. Sidky and B.D. Esry, Phys. Rev. Lett. \textbf{85}, 5086 (2000). \bibitem{Pindzola} M.S. Pindzola, F. Robicheaux, S.D. Loch, J.C. Berengut, T. Topcu, J. Colgan, M. Foster, D.C. Griffin, C.P. Ballance, D.R. Schultz, T. Minami, N.R. Badnell, M.C. Witthoeft, D.R. Plante, D.M. Mitnik, J.A. Ludlow and U. Kleiman, J. Phys. B: At. Mol. Phys. \textbf{40}, R39 (2007). \bibitem{Runge} E. Runge and E.K.U. Gross, Phys. Rev. Lett. \textbf{52}, 997 (1984). \bibitem{Marques} M.A.L. Marques and E.K.U. Gross, Annu. Rev. Phys. Chem. \textbf{55}, 427 (2004). \bibitem{Castro} A. Castro, M.A.L. Marques and A. Rubio, J. Chem. Phys. \textbf{121}, 3425 (2004). \bibitem{Lambert} J.D. Lambert, in \emph{Numerical methods for ordinary differential systems. The initial value problem} (Wiley, New York, 1991). \bibitem{Lapidus} L. Lapidus and J. H. Seinfeld, \emph{Numerical Solution of Ordinary Differential Equations} (Academic Press, New York and London, 1971). \bibitem{Fatunla1} S. O. Fatunla, Math. Comput. \textbf{32}, 1 (1978). \bibitem{Fatunla2} S. O. Fatunla, Math. Comput. \textbf{34}, 373 (1980). \bibitem{Smyth} E.S. Smyth, J.S. Parker and K.T. Taylor, Computer Physics Communication \textbf{114}, 1 (1998). \bibitem{Leforestier} C. Leforestier, R.H. Bisseling, C. Cerjan, M.D. Feit, R. Friesner, A. Guldberg, A. Hammerich, G. Jolicard, W. Karrlein, H.-D Meyer, N. Lipkin, O. Roncero and R. Kosloff, J. Comput. Phys. \textbf{94}, 59 (1991). \bibitem{Rotenberg} M. Rotenberg in \emph{Adv. in Atomic and Molecular Physics} edited by D.R. Bates and I. Estermann (Academic Press, 1970) vol. 6, p 233. \bibitem{Gottlieb} D. Gottlieb and S. Orzag, \emph{Numerical Analysis of Spectral Methods: Theory and Applications} (SIAM, Philadelphia, 1977). \bibitem{Eiglsperger} J. Eiglsperger, M. Sch\"onwetter, B. Piraux and J. Madro\~nero, Atomic Data and Nuclear Data Tables\textbf{98}, 120 (2012). \bibitem{Kato} T. Kato, Communications on Pure and Applied Mathematics \textbf{10}, 151 (1957). \bibitem{Bachau} H. Bachau, E. Cormier, P. Decleva, J.E. Hansen and F. Mart\'in, Reg. Prog. Phys.\textbf{64}, 1815 (2001). \bibitem{DiMenza} L. Di Menza, Numer. Funct. Anal. Optim. \textbf{18}, 759 (1997). \bibitem{Muga} J.G. Muga, J.P. Palao, B. Navarro and I.L. Egusquiza, Phys. Rep. \textbf{395}, 357 (2004). \bibitem{He} F. He, C. Ruiz and A. Becker, Phys. Rev. A \textbf{75}, 053407 (2007). \bibitem{Scrinzi1} A. Scrinzi, Phys. Rev. A \textbf{81}, 053845 (2010). \bibitem{Colgan} J. Colgan and M.S. Pindzola, Sur. Phys. J. D \textbf{66}, 284 (2012). \bibitem{Laulan} S. Laulan and H. Bachau, Phys. Rev. A \textbf{68}, 013409 (2003). \bibitem{Foumouo} E. Foumouo, G. Lagmago Kamta, G. Edah and B. Piraux, Phys. Rev. A \textbf{74}, 063409 (2006). \bibitem{Feist} J. Feist, S. Nagele, R. Pazourek, E. Persson, B.I. Schneider, L.A. Collins and J. Burgd\"orfer, Phys. Rev. A\textbf{77}, 043420 (2008). \bibitem{Malegat1} L. Malegat, H. Bachau, B. Piraux and F. Reynal, J. Phys. B \textbf{45}, 175601 (2012). \bibitem{Palacios} A. Palacios, T.N. Rescigno and C.W. McCurdy, Phys. Rev. A \textbf{79}, 033402 (2009). \bibitem{Malegat2} L. Malegat, H. Bachau, A. Hamido and B. Piraux, J. Phys. B \textbf{43}, 245601 (2010). \bibitem{Scrinzi2} A. Scrinzi, New J. Phys \textbf{14}, 085008 (2012). \bibitem{Hutchinson} S. Hutchinson, M.A. Lysaght and H. van der Hart, J. Phys.: Conf. Ser. \textbf{388}, 032022 (2012). \bibitem{Hamido} A. Hamido, J. Eiglsperger, J. Mado\~nero, F. Mota-Furtado, P. O'Mahony, A.L. Frapiccini and B. Piraux, Phys. Rev. A\textbf{84}, 013422 (2011). \bibitem{Madronero} J Madro\~{n}ero, B. Piraux, Phys Rev A \textbf{80}, 033409 (2009) \bibitem{Saad} Y. Saad , \emph{Iterative Methods for Sparse Linear Systems}, SIAM, 2000. \bibitem{Arnoldi} W.E. Arnoldi, Quart. Appl. Math 9 (1951), 17-29. \bibitem{Park} T.J. Park and J.C. Light, J Chem Phys \textbf{85}, 5870 (1986); see also E.S. Smyth, J.S. Parker and K.T. Taylor, Comp. Phys. Comm. \textbf{114},1 (1998). \bibitem{Saad2} Y. Saad, SIAM J. Numer. Anal. \textbf{29}, 209 (1992). \bibitem{Hochbruck} M. Hochbruck and Ch. Lubich, SIAM J. Numer. Anal. \textbf{34}, 1911 (1997). \bibitem{Magnus} W. Magnus, Comm. Pure and Appl. Math \textbf{VII}, 649 (1954). \bibitem{Iserles} A. Iserles and S.P. N\o rsett, Phil. Trans. Royal Society A \textbf{357}, 983 (1999). \bibitem{Houwen1} P. J. van der Houwen and B. P. Sommeijer, SIAM J. Sci. Stat. Comput. \textbf{12}, 1000 (1991). \bibitem{Houwen2} P.J. van der Houwen and B.P. Sommeijer, Appl. Numer. Math. \textbf{11}, 169 (1993). \bibitem{Vorst} H. van der Vorst, SIAM J. Sci. Statist. Comput. \textbf{13}, 631 (1992). \bibitem{Schroeter1} S. Schr\"oter, P.-A. Hervieux, G.~Manfredi, J.~Eiglsperger and J.~Madro\~nero, Phys. Rev. B\textbf{87}, 155413 (2013). \bibitem{Gorin} T. Gorin, T. Prosen, T.H. Seligman, and M. Znidaric, Phys. Rep. {\bf 435}, 33 (2006). \bibitem{Schroeter2} S. Schr\"oter, Dissertation, Technische Universit\"at M\"unchen (2013).\end{thebibliography} \end{document} %% }}\end{equation}}

13 — 1401.7226

\caption{(Color online) $U$ dependence of $\Gamma_d \overline P_d$ at $t'/t=0.2$ and $\rho=0.5$. The sign problem makes acquisition of low $T$ data increasingly difficult as $U$ increases. \textcolor{red}{Eliminate this figure? Why did we get lower in $T$ at $U=4$ than $U=2$?} \textcolor{cyan}{Laziness. Expect an updated version soon. However, if the observed tendency confirms, i.e., reaching a assymptotic value as $T\rightarrow0$, this might be a problem since we would never reach $``T_c"$. However, since this is only for the $8\times8$ case, there is still hope it will decrease for larger lattice sizes... Need to invest some time on this.} \textcolor{green}{keep this figure to show the dependence of U and explain why we didn't access $U=8$ which is argued optimal for pairing? Maybe the sign problem only becomes much worse for larger U.} \textcolor{magenta}{Waiting for new version to decide} \label{fig:tp0.2Ucomparison} %% FIG 12 }

14 — 1401.7520

\caption{\small Photon emission rates in plasmonic nanoparticle chains calculated by finite-difference time-domain simulations. A dipole source is placed inside a particular sphere with a small hole of negligible size, acting as an emitter. The emission rate is defined by Eq.~(\ref{eq:emission}), which represents approximately the local density of states (LDOS). (a) Emission rate in single sphere, diatomic chain, and connected chain. The last one reveals the existence of an edge state. Videos of corresponding time-domain fields are attached, in which longitudinal component of electric fields are shown. \textcolor{blue}{Media~1}: connected chain; \textcolor{blue}{Media~2}: diatomic chain; \textcolor{blue}{Media~3}: monatomic chain; \textcolor{blue}{Media~4}: single sphere . (b) Emission rate for connected case with different $t_L$. $d/2 = 75$ nm and $t_R = d-t_L$.}

15 — 1401.8174

\caption{Truncated \highlight{disks} -- a construction of $2$ components in $\mathbb{E}^2$ without integral distances.}

\caption{Truncated \highlight{discs} -- arranged on a parabola.}

16 — 1401.8199

\caption{Simulation with IEC wind gust. \textcolor{blue}{Blue}: True states from FAST simulation; \textcolor{red}{red}: Estimated states; Initial values: $\theta_{s,0} = 0\,\text{rad}$, $\hat{\theta}_{s,0} = 0.1\,\text{rad}$, $\omega_{r,0} = \omega_{g,0} = 1.267 \,\frac{\text{rad}}{\text{s}}$, $\hat{\omega}_{r,0} = \hat{\omega}_{g,0} = 0 \,\frac{\text{rad}}{\text{s}}$, $v_0 = 18 \,\frac{\text{m}}{\text{s}}$, $\hat{v}_0 = 1 \,\frac{\text{m}}{\text{s}}$. {The torsion angle is not directly available from the FAST outputs and was obtained by integrating the speed error signal from the FAST outputs of rotor and generator speed (corrected by the gear ratio). The pitch angle is only shown for reference.}}

\caption{Turbulent wind simulation results. \textcolor{blue}{Blue}: True states from FAST simulation; \textcolor{red}{red}: Estimated states; Initial values: $\theta_{s,0} = 0\,\text{rad}$, $\hat{\theta}_{s,0} = 0.1\,\text{rad}$, $\omega_{r,0} = \omega_{g,0} = 1.267 \,\frac{\text{rad}}{\text{s}}$, $\hat{\omega}_{r,0} = \hat{\omega}_{g,0} = 0 \,\frac{\text{rad}}{\text{s}}$, $\hat{v}_0 = 1 \,\frac{\text{m}}{\text{s}}$. The \textcolor{blue}{blue} wind speed signal from FAST is the nominal downwind component of the hub-height wind speed, not the rotor effective wind speed.}

17 — 1401.8234

\caption{(color online) Strain response $\gamma(t)$ in a 4\%~wt. casein gel acidified with 1\%~wt. GDL for an imposed shear stress $\sigma=200$ (\textcolor{red!25!black}{$\bullet$}), 300 (\textcolor{red!50!black}{$\blacktriangledown$}), 400 (\textcolor{red!75!black}{$\blacksquare$}), 550 (\textcolor{red}{$\blacktriangle$}) and 1000~Pa (\textcolor{orange!50!red}{$\blacklozenge$}) from right to left. The gap width is 1~mm. Gray dashes show $\gamma=1$. Inset: failure time $\tau_f$ vs $\sigma$. The red line is the best power-law fit $\tau_f=A\sigma^{-\beta}$ with $\beta=5.45\pm 0.05$ and $A=(4.2\pm 0.1)\,10^{17}$~s.Pa$^\beta$. \label{fig1}}

\caption{(color online) (a)~Local velocity $\langle v(r,z,t)\rangle_z$ and (b)~local strain field $\langle\gl(r,z,t)\rangle_z$ averaged over the vertical direction $z$ at various times during primary creep: $t/\tau_f=1.9\,10^{-3}$ ($\bullet$), $1.7\,10^{-2}$ (\textcolor{blue}{$\blacktriangledown$}) and 0.15 (\textcolor{red}{$\blacksquare$}). Solid lines are linear profiles. The arrows in (a) indicate the velocity of the inner cylinder inferred from the current shear rate. (c)~Spatiotemporal diagram of the local velocity $\langle v(r,z,t)\rangle_r$ averaged over the radial direction $r$ and plotted in linear color levels as a function of $z$ and $t/\tau_f$. (d)~Standard deviation $\delta_z v(t)$ of $\langle v(r,z,t)\rangle_r$ taken over the vertical direction $z$ (thick black line) together with corresponding standard deviation $\delta_r v(t)$ computed over the radial direction $r$ on the $z$-average $\langle v(r,z,t)\rangle_z$ (thin red line). (e)~Fracture length $\ell(t)$ vs $(\tau_f-t)/\tau_f$ as inferred from direct visualization ($\bullet$, average over 6 different fractures, error bars show the standard deviation) and from ultrasonic imaging ($\circ$) and normalized by the height $H$ of the TC cell. Gray dots show the visualization data for the longest fracture which leads to the failure of the sample at $\tau_f$. Red lines are the best fits $\ell(t)=a+b\log(1-t/\tau_f)$ to the visualization data. Same experiment as in Fig.~\ref{fig3} and Supplemental Movie~2. \label{fig4}}

18 — 1402.0096

\caption{Best matches (in \textcolor{red}{red}) corresponding to a fixed patch (depicted in \textcolor{green}{green}) \newbf{in an image with lot of self-similarities.}}

19 — 1402.0383

\caption{Luminosity in 3 \gray\bands for a star-forming disk galaxy with ($R_{\textrm{max}}$, $z_{\textrm{max}}$)=(10\,kpc, 2\,kpc), 4 profiles of molecular gas (given in abscissa), and 7 molecular gas densities for each profile ($n_{\textrm{H}_2}=$5, 10, 20, 50, 100, 200, 500\,H$_2$\dunit). To illustrate the effects of the gas distribution, the \gray\luminosities are given for the same input cosmic-ray luminosity.}

\caption{Distribution of the luminosity in 3 \gray\bands, in terms of physical processes and emitting particles, for a star-forming disk galaxy with ($R_{\textrm{max}}$, $z_{\textrm{max}}$)=(10\,kpc, 2\,kpc), 4 profiles of molecular gas (given in abscissa), and 7 molecular gas densities for each profile ($n_{\textrm{H}_2}=$5, 10, 20, 50, 100, 200, 500\,H$_2$\dunit).}

\caption{Spectra of pion decay (top), Bremsstrahlung (middle), and inverse-Compton (bottom) emission for a star-forming disk galaxy with ($R_{\textrm{max}}$, $z_{\textrm{max}}$)=(10\,kpc, 2\,kpc) and a 4-6\,kpc ring distribution of molecular gas. The solid, dashed, and dot-dashed curves correspond to$n_{\textrm{H}_2}$=5, 50, and 500\,H$_2$\dunit, respectively. For the leptonic processes, the contribution from primary electrons and secondary positrons are shown. In all plots, the total \gray\emission from all processes and all particles is shown in black.}

\caption{Photon spectral indices over 3 \gray\bands for a star-forming disk galaxy with ($R_{\textrm{max}}$, $z_{\textrm{max}}$)=(10\,kpc, 2\,kpc), 4 profiles of molecular gas, and 7 molecular gas densities for each profile ($n_{\textrm{H}_2}=$5, 10, 20, 50, 100, 200, 500\,H$_2$\dunit). Blue, red, green, and purple curves correspond to the 500\,pc core, 1\,kpc core, 2-3\,kpc ring, and 4-6\,kpc ring gas profiles, respectively. Note that the energy bands differ from those used for the luminosities.}

\caption{Luminosity in 3 \gray\bands as a function of star formation rate and cosmic-ray power, for a star-forming disk galaxy with ($R_{\textrm{max}}$, $z_{\textrm{max}}$)=(10\,kpc, 2\,kpc), 4 profiles of molecular gas (given in abscissa), and 7 molecular gas densities for each profile ($n_{\textrm{H}_2}=$5, 10, 20, 50, 100, 200, 500\,H$_2$\dunit). The cosmic-ray input power generated by the galaxy is assumed to scale with total star formation rate, using the Milky Way as normalisation.}

\caption{Luminosity in the 100\mev-100\gev\band as a function of total infrared luminosity and star formation rate for the complete sample of models. The orange region is the uncertainty range of the correlation determined experimentally. The top panel shows the result for the basic model, and the middle panel shows the result obtained with larger halos for the small galaxies and using a correction for starlight leakage. The bottom panel shows the ratio of 100\mev-100\gev\to total infrared luminosities.}

\caption{Radio luminosity at 1.4\,GHz as a function of far-infrared luminosity for the complete sample of models. The orange region is the uncertainty range of the linear scaling derived from the observed far-infrared - radio correlation. The top panel shows the result for the basic model, and the middle panel shows the result obtained with larger halos for the small galaxies and using a correction for starlight leakage. The bottom panel shows the ratio of 1.4\,GHz to far-infrared luminosities.}

20 — 1402.1181

\caption{$\DC$ profile of the semileptonic asymmetries $\asld$ and $\asls$; the {\color{blue} blue} line corresponds to the NP scenario -- \eq{eq:33NP:Mix01} --, the {\color{red} red dashed} line corresponds to the SM case. Notice that for $\asls$ the SM range is too narrow to be resolved on this scale.\label{fig:UnNP:01}}

\caption{$\DC$ profile of $\aslb$; the {\color{blue} blue} line corresponds to the NP scenario, the {\color{red} red dashed} line corresponds to the SM case. The last D0 measurements gives $\aslb=(-4.96\pm 1.69)\cdot 10^{-3}$ \cite{Abazov:2013uma}.\label{fig:UnNP:04}}

\caption{$\DC$ profile of the combinations of semileptonic asymmetries $\asls\pm\asld$; the {\color{blue} blue} lines correspond to the NP scenario -- \eq{eq:33NP:Mix01} --, the {\color{red} red dashed} lines correspond to the SM case.\label{fig:UnNP:05}}

\caption{$\DC$ 68\%, 95\% and 99\% CL regions. {\color{blue} Blue} regions correspond to the NP scenario, {\color{red} red} regions correspond to the SM case. Notice that with the scales in fig. \ref{fig:UnNP:06b}, the SM region is barely a point.\label{fig:UnNP:06}}

\caption{$\DC$ profiles of semileptonic asymmetries $\aslq$; the {\color{blue} blue} lines correspond to the $4\times 4$ unitary NP scenario -- eqs. \refeq{eq:Un44Mixq:01} and \refeq{eq:Un44Mixq:02} --, the {\color{red} red dotted} lines correspond to the $3\times 3$ unitary NP scenario of section \ref{SEC:3x3NP}, the {\color{red} red dashed} lines correspond to the SM case. The last D0 measurement gives $\aslb=(-4.96\pm 1.69)\cdot 10^{-3}$ \cite{Abazov:2013uma}.\label{fig:NoUnNP:01}}

\caption{$\DC$ 68\%, 95\% and 99\% CL regions. {\color{blue} Blue} regions correspond to the $4\times 4$ unitary NP scenario, {\color{red} red} regions correspond to the SM case.\label{fig:NoUnNP:02}}

21 — 1402.1450

\caption{Running times of model checking the extinction formula in (\ref{eq:ext}) for the SIR model. The experiments have been performed in an Intel\textregistered\Xeon\texttrademark\E5410 @ 2.33GHz PC running Linux.}

\caption{Running times of model checking the variation formula in (\ref{eq:variationFormula}) for the LacZ model. The experiments have been performed in an Intel\textregistered\Xeon\texttrademark\E5410 @ 2.33GHz PC running Linux.}

\caption{Running times of model checking the expression burst formula in (\ref{eq:burstFormula}) for the LacZ model. The experiments have been performed in an Intel\textregistered\Xeon\texttrademark\E5410 @ 2.33GHz PC running Linux.}

22 — 1402.1650

\caption{\textcolor{blue}{Spectrum of the microwave signal generated by\,$\sim4.6$\,GHz repetition rate pulses impinging on a KTP crystal placed inside a coaxial waveguide as shown in Fig.~\ref{sc}(b). The spectrum extends up to the sampling frequency of the oscilloscope (LeCroy model SDA 820Zi-A, 20\,GHz oscilloscope).} }

23 — 1402.2525

\caption{ COMPTEL error location contours for the most significant COMPTEL detection of GRO~J1823-12, obtained in the 10-30 MeV band for the sum of all data. The contour lines are plotted on a map in galactic coordinates (l, b) of the LS~5039-region. The error contours start with 1\sig\with steps of 1\sig. The sky positions (circles) of all \gray\source are shown, which are listed in the second Fermi catalog and which are within a search radius of 2\deg\around the pixel center of the best-fit source location (l/b:17.5\deg/-0.5\deg). }

\caption{ The time-averaged COMPTEL energy spectrum of GRO~J1823-12 for the sum of all data (see Fig.~\ref{ls5039_specs}) is compared to the extrapolations of the Fermi/LAT spectrum of the 6 closest \gray\sources, and to the extrapolation of the time-averaged X-ray spectrum (1-10~keV, dashed blue line) of LS~5039 as measured by{\it Suzaku} (Takahasi et al. 2009). The thick solid lines represent the spectral shape derived from best-fit Fermi/LAT integral fluxes between 100 and 300~MeV by assuming a power-law shape of fixed index. This index was derived by a power-law throughout the Fermi/LAT energy band. The thin solid lines represent the spectral extrapolations down to 1~MeV. For the known Fermi/LAT sources, the spectral extrapolation for LS~5039 is coming closest to the COMPTEL 10-30~MeV measurement. Also the power-law extrapolation of the {\it Suzaku}-measured X-ray spectrum is in reasonable agreement with the COMPTEL spectrum of GRO~J1823-12. For more details see text. }

\caption{ The orbital light curve of LS~5039 in the 10-30~MeV COMPTEL band for the sum of all data. The lightcurve is folded with the orbital period of \sm3.9 days and given in phase bins of 0.2. The two broader phase periods, defined as {\it INFC} and {\it SUPC} are indicated. A flux increase during the {\it INFC} period is obvious. In the phase bin containing the inferior conjunction the source is roughly three times brighter than in the phase bin containing the superior conjunction. In general the 10-30~MeV \gray\lightcurve is consistent in phase and amplitude with the one at TeV\grays. }

\caption{ The X-ray to TeV \gray\SED of LS~5039. The COMPTEL soft\gray\spectra (sum of all data) for the{\it INFC} and {\it SUPC} orbital phases are combined with similar spectra from the X-ray \citep[{\it Suzaku},][]{Takahashi09}, the MeV/GeV \citep[Fermi/LAT,][]{Hadasch12}, and the TeV band \citep[HESS,][]{Aharonian06a} to build a high-energy SED of LS~5039. The SED shows that 1) the emission maximum at energies above 1~keV and 2) a switch in radiation dominance is occuring a MeV energies. We like to note that the lines -- solid and dashed -- in the X-ray spectra represent the model calculations of \citet{Takahashi09} on the emission pattern of LS~5039. They do not present a fit to their measured X-ray spectra. }

24 — 1402.2705

\caption{\label{fig:elliptical}\textcolor{gray}({Color online) (a) Energy branches as a function of the electric field in the region of the $s$-$d$ resonances at $x_\mathrm{s}=1.8$~nm with $y_\mathrm{s}=0$. (b) Energy branches as a function of electric field in the region of the $s$-$d$ resonances in the case of elliptical dots.} }

25 — 1402.2730

\caption{(A) $f_{IR}(\tau)$ for $\gamma_{o} = 0.05$ ({\color{Brown} $\boldsymbol \circ$}), $\gamma_{o} = 0.12$ ({\color{Green} $\boldsymbol \bigstar$}), $\gamma_{o} = 0.215$ ({\color{Blue} $\boldsymbol \diamond$}), $\gamma_{o} = 0.25$ ({\color{Red} $\boldsymbol \blacksquare$}) and $\gamma_{o} = 0.37$ ({\color{Black} $\boldsymbol \triangle$}). The dashed lines are linear fits to the data. (B) $f_{IR}^{\infty}$ as a function of $\gamma_{o}$. Inset to B shows $f_{IR}^{\infty}$ versus $|\gamma_{o} - \gamma_{c}^{Mi}|$. $f_{IR}^{\infty}(\gamma_{o}) = f_{IR}^{ss}(\gamma_{o}) - f_{IR}(\gamma_{o} = 0) $, where $f_{IR}^{ss}$ is the steady state fraction of irreversible rearrangements obtained by averaging over the shaded region in A. $f_{IR}(\gamma_{o} = 0)$ = 0.023. The red curve is a power-law fit to the data. $\gamma_{c}^{Mi}$ and $\beta$ were extracted by minimizing $\chi^2$. (C) $G''(\tau)$ for $\gamma_{o} = 0.04$ ({\color{Blue} $\boldsymbol \bullet$}), $\gamma_{o} = 0.25$ ({\color{Red} $\boldsymbol \blacksquare$}) and $\gamma_{o} = 0.45$ ({\color{Black} $\boldsymbol \triangledown$}). The curves are best fits to the data. (D) Relaxation time $\tau _{s}$ versus $\gamma_{o}$. The solid curves represent power-law fits to the data. Inset to D shows $\tau _{s}$ versus $|\gamma_{o} - \gamma_{c}^{Rh}|$ . The red line is a fit to the data whereas the blue line is a guide to the eye and has the same slope as the red line. ({\color{Blue} $\boldsymbol \bullet$}) and ({\color{Red} $\boldsymbol \square$}) correspond to the regimes $\gamma_{o} < 0.25$ and $\gamma_{o} > 0.25$, respectively. Black Diamonds ({\color{Black} $\boldsymbol \diamond$}) correspond to $\tau _{s}$ obtained from independent measurements on the same sample.}

\caption{(A) Color map of $u_{i}$ for $\tau = 2$. The solid spheres represent irreversible particles at the end of the same cycle. Inset shows the distribution of $u_{i}$. (B)-(C) Representative snapshots with the top 10$\%$ high $u_{i}$ particles shown as big solid spheres and the remaining shown as small circles. The colors are a visual aid to help demarcate clusters. (D)-(F) Distribution of cluster size $P(n)$. In (D)-(F) P(n) for $\tau = 2$ to $10$ ({$\boldsymbol \bullet$}), for $\tau = 20$ to $30$ ({\color{Red} $\boldsymbol \circ$}) and for $\tau = 37$ to $47$ ({\color{Green} $\boldsymbol \blacktriangle$}). (G) $\langle n \rangle$ as a function of $\gamma_{o}$ is shown as ({\color{Red} $\boldsymbol \bullet$}). $G''$ from bulk rheology is shown as ({$\boldsymbol \square$}).}

26 — 1402.2958

\caption{\label{fig:hyst_explanation} Origin of hysteresis. {\bf a}, A schematic of the energy landscape of a hysteretic system. As a function of an order parameter $\kappa$, the energy can have local minima (\textcolor{darkred}{$\blacksquare$},\textcolor{darkblue}{\CIRCLE}), which represent stable states, separated by a local maximum (\textcolor{darkgreen}{$\bigstar$}), which forms an energy barrier $E_b$. This landscape is shown for five values of the applied field $F$ (for superfluidity, $F\rightarrow\Omega$, the rotation rate of the trap). {\bf b}, Plotted as a function of $\Omega$ for a superfluid, the energy of the minima (solid) and maximum (dashed) form a swallowtail (upper), which exhibits hysteresis (lower). {\bf c}, This swallowtail structure is periodic in $\Omega_0$; states above $E_2$ are unstable.}

\caption{\label{fig:critical_velocity} Extracted critical velocities vs. $U_2$. The red circles (blue squares) show the critical velocities extracted from $\Omega_c^+$ ($\Omega_c^-$). The magenta line and band show the estimate and uncertainty of the local speed of sound. The green, dashed line and band show the best fit of the toy model of vortex creation and its statistical uncertainty. All uncertainties are $1\sigma$. {\bf a}, A diagram of a vortex/anti-vortex pair in a weak link of width $d$ and a vortex/anti-vortex separation $s$. {\bf b}, Energy landscape as a function of $s$ for three different values of the velocity in the weak link region $v_m$, showing the stable states (\textcolor{darkred}{$\blacksquare$},\textcolor{darkblue}{\CIRCLE}) and the energy barrier (\textcolor{darkgreen}{$\bigstar$}).}

27 — 1402.3401

\caption{Top-5 hosts for \policyredirect requests in \df.}

28 — 1402.3514

\caption{$\mbox{bias}(\pmb V_q)$ for $p=100$, shift (top) and point-mass (bottom) as a function of $\nu$. \textcolor{ROBPCA}{\underline{ROBPCA}}, \textcolor{PcaPP}{\underline{PcaPP}}, \textcolor{PcaL}{\underline{PcaL}}, \textcolor{HCS}{\underline{FastHCS}}.}

\caption{$\mbox{bias}(\pmb V_q)$ for $p=400$, shift (top) and point-mass (bottom) as a function of $\nu$. \textcolor{ROBPCA}{\underline{ROBPCA}}, \textcolor{PcaPP}{\underline{PcaPP}}, \textcolor{PcaL}{\underline{PcaL}}, \textcolor{HCS}{\underline{FastHCS}}.}

29 — 1402.4252

\caption{{\bf (a)} the steady state of the equation with $W(\mathbf{x}) =|\mathbf{x}|^2/2 - \ln |\mathbf{x}|$; {\bf (b)} the steady state with the same $W(\mathbf{x})$, regularized by quadratic diffusion $\nabla\cdot\big(\rho\nabla(\epsilon\rho)\big)$. The exact steady state without diffusion is the characteristic function of the unit disk with density $\frac{1}{\pi}$.} \label{fig:quadlog2d} \end{figure} \end{example} \begin{example}[{\bf Steady mill solutions}]\label{ex311} Another common pattern observed for the self-propelled particle systems with an attractive-repulsive kernel in 2-D is the rotating mill~\cite{MR2507454}, and the steady pattern can be obtained from the equation \[ \rho_t = \nabla \cdot \big(\rho \nabla(W\ast \rho-\frac{\alpha}{\beta} \log |\mathbf{x}|)\big), \quad \mathbf{x} \in \mathbb{R}^2, \] with some positive constants $\alpha$ and $\beta$. For the kernel $W(\mathbf{x}) = \frac{1}{2}|\mathbf{x}|^2-\ln |\mathbf{x}|$, the steady state is still a constant $\rho_\infty=2$ on an annulus, whose inner and outer radius are given by $$ R_0 = \sqrt{\frac{\alpha}{\beta}},\quad R_1 = \sqrt{\frac{\alpha}{\beta}+\frac{M}{2\pi}}, $$ with the total conserved mass $M=\int_{\mathbb{R}^d} \rho d{\bf x}$. For other more realistic kernels like the Morse type~\cite{MR2507454} or Quasi-Morse type~\cite{Carrillo2013112}, the radial density is in general more concentrated near the inner radius, but the explicit form of $\rho_\infty$ can not be obtained in general. Numerical diffusion, in the form of $\epsilon \nabla \cdot(\rho \nabla \rho)$, is still needed to prevent the overshoot and the resulting steady states with $\epsilon= 0.2((\Delta x)^2+(\Delta y)^2)$ are shown in Figure~\ref{fig:mill} for two different potentials. \begin{figure}[thp] \centering \subfloat[$W(\mathbf{x})=|\mathbf{x}|^2/2-\ln |\mathbf{x}|$]{\includegraphics[totalheight=0.29\textheight]{Logquadmill}} $~~$ \subfloat[$W(\mathbf{x})=\lambda\big(V(|\mathbf{x}|)-CV(|\mathbf{x}|/\ell)\big)$]{\includegraphics[totalheight=0.29\textheight]{QMmill}} \caption{The steady density $\rho_\infty$ for the rotating mill with $\Delta x=\Delta y=0.05$. {\bf (a)} $\alpha = 0.25$, $\beta=2\pi$; {\bf (b)} $V(r) = -K_0(kr)/2\pi$, where $K_0(r)$ is the modified Bessel function of the second kind and the parameters $C = 10/9, \ell=0.75, k =0.5, \lambda=100, \alpha = 1.0, \beta=40$ are taken from\cite{Carrillo2013112}.} \label{fig:mill} \end{figure} \end{example} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\section{Conclusion}\label{sec4} \begin{acknowledgments} JAC acknowledges support from projects MTM2011-27739-C04-02, 2009-SGR-345 from Ag\`encia de Gesti\'o d'Ajuts Universitaris i de Recerca-Generalitat de Catalunya, and the Royal Society through a Wolfson Research Merit Award. JAC and YH were supported by Engineering and Physical Sciences Research Council (UK) grant number EP/K008404/1. The work of AC was supported in part by the NSF Grant DMS-1115682. The authors also acknowledge the support by NSF RNMS grant DMS-1107444.\end{acknowledgments} \begin{thebibliography}{10} \bibitem{MR2401600}{\sc L.~Ambrosio, N.~Gigli, and G.~Savar{\'e}}, {\em Gradient flows in metric spaces and in the space of probability measures}, Lectures in Mathematics ETH Z\"urich, Birkh\"auser Verlag, Basel, second~ed., 2008.\bibitem{BCLR2}{\sc D.~Balagu{\'e}, J.~A. Carrillo, T.~Laurent, and G.~Raoul}, {\em Dimensionality of local minimizers of the interaction energy}, Arch. Ration. Mech. Anal., 209 (2013), pp.~1055--1088. \bibitem{BCCP}{\sc D.~Benedetto, E.~Caglioti, J.~A. Carrillo, and M.~Pulvirenti}, {\em A non-{M}axwellian steady distribution for one-dimensional granular media}, J. Statist. Phys., 91 (1998), pp.~979--990. \bibitem{MR1471181}{\sc D.~Benedetto, E.~Caglioti, and M.~Pulvirenti}, {\em A kinetic equation for granular media}, RAIRO Mod\'el. Math. Anal. Num\'er., 31 (1997), pp.~615--641.\bibitem{Filbet}{\sc M.~Bessemoulin-Chatard and F.~Filbet}, {\em A finite volume scheme for nonlinear degenerate parabolic equations}, SIAM J. Sci. Comput., 34 (2012), pp.~B559--B583. \bibitem{BCC}{\sc A.~Blanchet, E.~A. Carlen, and J.~A. Carrillo}, {\em Functional inequalities, thick tails and asymptotics for the critical mass {P}atlak-{K}eller-{S}egel model}, J. Funct. Anal., 262 (2012), pp.~2142--2230. \bibitem{BCL}{\sc A.~Blanchet, J.~A. Carrillo, and P.~Lauren{\c{c}}ot}, {\em Critical mass for a {P}atlak-{K}eller-{S}egel model with degenerate diffusion in higher dimensions}, Calc. Var. Partial Differential Equations, 35 (2009), pp.~133--168. \bibitem{BCM}{\sc A.~Blanchet, J.~A. Carrillo, and N.~Masmoudi}, {\em Infinite time aggregation for the critical {P}atlak-{K}eller-{S}egel model in {$\mathbb{R}^2$}}, Comm. Pure Appl. Math., 61 (2008), pp.~1449--1481. \bibitem{BCW}{\sc M.~Burger, J.~A. Carrillo, and M.-T. Wolfram}, {\em A mixed finite element method for nonlinear diffusion equations}, Kinet. Relat. Models, 3 (2010), pp.~59--83. \bibitem{BDF}{\sc M.~Burger, M.~di~Francesco, and M.~Franek}, {\em Stationary states of quadratic diffusion equations with long-range attraction}, Commun. Math. Sci., 11 (2013), pp.~709--738. \bibitem{BFH}{\sc M.~Burger, R.~Fetecau, and Y.~Huang}, {\em Stationary states and asymptotic behaviour of aggregation models with nonlinear local repulsion}, 2013. \bibitem{CC}{\sc V.~Calvez and J.~A. Carrillo}, {\em Volume effects in the {K}eller-{S}egel model: energy estimates preventing blow-up}, J. Math. Pures Appl. (9), 86 (2006), pp.~155--175. \bibitem{CD}{\sc J.~F. Campos and J.~Dolbeault}, {\em Asymptotic estimates for the parabolic-elliptic keller-segel model in the plane}, preprint, (2013). \bibitem{Carrillo2013112}{\sc J.~Carrillo, S.~Martin, and V.~Panferov}, {\em A new interaction potential for swarming models}, Physica D: Nonlinear Phenomena, 260 (2013), pp.~112 -- 126. \bibitem{MR2507454}{\sc J.~A. Carrillo, M.~R. D'Orsogna, and V.~Panferov}, {\em Double milling in self-propelled swarms from kinetic theory}, Kinet. Relat. Models, 2 (2009), pp.~363--378. \bibitem{CFP}{\sc J.~A. Carrillo, L.~C.~F. Ferreira, and J.~C. Precioso}, {\em A mass-transportation approach to a one dimensional fluid mechanics model with nonlocal velocity}, Adv. Math., 231 (2012), pp.~306--327. \bibitem{cmcv-03}{\sc J.~A. Carrillo, R.~J. McCann, and C.~Villani}, {\em Kinetic equilibration rates for granular media and related equations: entropy dissipation and mass transportation estimates}, Rev. Mat. Iberoam., 19 (2003), pp.~971--1018. \bibitem{cmcv-06} \leavevmode\vrule height 2pt depth -1.6pt width 23pt, {\em Contractions in the 2-{W}asserstein length space and thermalization of granular media}, Arch. Ration. Mech. Anal., 179 (2006), pp.~217--263. \bibitem{CT00}{\sc J.~A. Carrillo and G.~Toscani}, {\em Asymptotic {$L^1$}-decay of solutions of the porous medium equation to self-similarity}, Indiana Univ. Math. J., 49 (2000), pp.~113--142. \bibitem{d2006self}{\sc M.~R. D'Orsogna, Y.-L. Chuang, A.~L. Bertozzi, and L.~S. Chayes}, {\em Self-propelled particles with soft-core interactions: patterns, stability, and collapse}, Phys. Rev. Lett., 96 (2006), p.~104302. \bibitem{FR11}{\sc K.~Fellner and G.~Raoul}, {\em Stable stationary states of non-local interaction equations}, Math. Models Methods Appl. Sci., 20 (2010), pp.~2267--2291. \bibitem{FR211} \leavevmode\vrule height 2pt depth -1.6pt width 23pt, {\em Stability of stationary states of non-local equations with singular interaction potentials}, Math. Comput. Modelling, 53 (2011), pp.~1436--1450. \bibitem{FHK}{\sc R.~C. Fetecau, Y.~Huang, and T.~Kolokolnikov}, {\em Swarm dynamics and equilibria for a nonlocal aggregation model}, Nonlinearity, 24 (2011), pp.~2681--2716. \bibitem{GST}{\sc S.~Gottlieb, C.-W. Shu, and E.~Tadmor}, {\em Strong stability-preserving high-order time discretization methods}, SIAM Rev., 43 (2001), pp.~89--112. \bibitem{keller1970initiation}{\sc E.~F. Keller and L.~A. Segel}, {\em Initiation of slime mold aggregation viewed as an instability}, Journal of Theoretical Biology, 26 (1970), pp.~399--415. \bibitem{PhysRevE.63.017101}{\sc H.~Levine, W.-J. Rappel, and I.~Cohen}, {\em Self-organization in systems of self-propelled particles}, Phys. Rev. E, 63 (2000), p.~017101. \bibitem{LT04}{\sc H.~Li and G.~Toscani}, {\em Long-time asymptotics of kinetic models of granular flows}, Arch. Ration. Mech. Anal., 172 (2004), pp.~407--428. \bibitem{LN}{\sc K.-A. Lie and S.~Noelle}, {\em On the artificial compression method for second-order nonoscillatory central difference schemes for systems of conservation laws}, SIAM J. Sci. Comput., 24 (2003), pp.~1157--1174. \bibitem{MR1451422}{\sc R.~J. McCann}, {\em A convexity principle for interacting gases}, Adv. Math., 128 (1997), pp.~153--179. \bibitem{NT}{\sc H.~Nessyahu and E.~Tadmor}, {\em Nonoscillatory central differencing for hyperbolic conservation laws}, J. Comput. Phys., 87 (1990), pp.~408--463. \bibitem{MR1842429}{\sc F.~Otto}, {\em The geometry of dissipative evolution equations: the porous medium equation}, Comm. Partial Differential Equations, 26 (2001), pp.~101--174. \bibitem{MR1485778}{\sc E.~B. Saff and V.~Totik}, {\em Logarithmic potentials with external fields}, vol.~316 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Springer-Verlag, Berlin, 1997. \newblock Appendix B by Thomas Bloom. \bibitem{St}{\sc G.~Str{\"o}hmer}, {\em Stationary states and moving planes}, in Parabolic and {N}avier-{S}tokes equations. {P}art 2, vol.~81 of Banach Center Publ., Polish Acad. Sci. Inst. Math., Warsaw, 2008, pp.~501--513. \bibitem{Swe}{\sc P.~Sweby}, {\em High resolution schemes using flux limiters for hyperbolic conservation laws}, SIAM J. Numer. Anal., 21 (1984), pp.~995--1011. \bibitem{MR2257718}{\sc C.~M. Topaz, A.~L. Bertozzi, and M.~A. Lewis}, {\em A nonlocal continuum model for biological aggregation}, Bull. Math. Biol., 68 (2006), pp.~1601--1623. \bibitem{MR1812737}{\sc G.~Toscani}, {\em One-dimensional kinetic models of granular flows}, M2AN Math. Model. Numer. Anal., 34 (2000), pp.~1277--1291. \bibitem{vLeV}{\sc B.~van Leer}, {\em Towards the ultimate conservative difference scheme. {V}. {A} second-order sequel to {G}odunov's method}, J. Comput. Phys., 32 (1979), pp.~101--136. \bibitem{Vaz}{\sc J.~L. V{\'a}zquez}, {\em The porous medium equation}, Oxford Mathematical Monographs, The Clarendon Press Oxford University Press, Oxford, 2007. \newblock Mathematical theory. \bibitem{V}{\sc J.~J.~L. Vel{\'a}zquez}, {\em Point dynamics in a singular limit of the {K}eller-{S}egel model. {I}. {M}otion of the concentration regions}, SIAM J. Appl. Math., 64 (2004), pp.~1198--1223. \bibitem{MR1964483}{\sc C.~Villani}, {\em Topics in optimal transportation}, vol.~58 of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, 2003. \bibitem{MR2001757}{\sc J.~von~zur Gathen and J.~Gerhard}, {\em Modern computer algebra}, Cambridge University Press, Cambridge, second~ed., 2003. \bibitem{YaoBer}{\sc Y.~Yao and A.~L. Bertozzi}, {\em Blow-up dynamics for the aggregation equation with degenerate diffusion}, Physica D: Nonlinear Phenomena, 260 (2013), pp.~77 -- 89. \end{thebibliography} \end{document} }

30 — 1402.4288

\caption{(Color online) Dynamics of the DNP process via the Overhauser effect when pumping the \textit{b-c} transition. Log $1-P(t)$ is plotted as a function of pumping time $t_p$ for 0.20 K - \textbf{\textbullet}, 0.27 K - {\color{red} $\blacksquare$ } and 0.92 K - {\color{blue} $\blacktriangle$}. The lines are guides for the eye.}

\caption{(Color online) Nuclear relaxation measured with different starting polarization values at 1.37 K. \textbf{\textbullet} - relaxation curve after starting with $P_{0}\approx 1$, {\color{green}$\blacksquare$} and {\color{blue}$\blacktriangle$} - relaxation curves with lower starting values of $P_{0}\approx 0.6$ and $P_{0}\approx 0.3$. The solid line is bi-exponential fit to the data. The dashed lines are exponential fits to the initial parts of the data. The data for $P_{0}\approx 0.6$ and $P_{0}\approx 0.3$ are shifted in time to match the starting points with the $P_{0}\approx 1$ curve. The inset shows full time range of the relaxations.}

\caption{(Color online) Nuclear relaxation rates {\color{red}$\blacksquare$} - $1/T^{'}_{ab}$ and \textbf{\textbullet} - $1/T^{''}_{ab}$ as a function of inverse temperature and exponential fits to the data. Data of van Tol \emph{et al.} \cite{vanTol} are given as {\color{magenta}$\square$} for comparison.}

31 — 1402.5295

\caption{Real parts of \textcolor{red}{$\zeta(\frac{1}{2}+\myi t)$ (dashed)} and \textcolor{blue}{$\Delta_{17}(\frac{1}{2}+\myi t)$ (dotted). }}

\caption{Imaginary parts of \textcolor{red}{$\zeta(\frac{1}{2}+\myi t)$ (dashed)} and \textcolor{blue}{$\Delta_{17}(\frac{1}{2}+\myi t)$ (dotted). }}

\caption{Coefficients $\delta_{101,n}$ \textcolor{red}{(disks} for even $n$ and \textcolor{blue}{triangles} for odd $n$)}

\caption{Coefficients $\delta_{\bigN,n}$ (\textcolor{red}{disks} for even $n$ and \textcolor{blue}{triangles} for odd $n$)}

\caption{Coefficients $\delta_{\bigN,n}$ (\textcolor{red}{disks} for even $n$ and \textcolor{blue}{triangles} for odd $n$)}

32 — 1402.5410

\caption{(a) Sketch of the studied Ta/\textcolor{blue2}{Pt}/\textcolor{orange}{Co}/\textcolor{red}{Ir}($t_{\mathrm{Ir}}$)/\textcolor{blue2}{Pt} layer stacks with a varying Ir thickness. (b) Polar Kerr hysteresis loops for samples with various Ir thickness. (c) Anisotropy field $\mu_0 H_\mathrm{K}$ and areal magnetization $M_\mathrm{s}t$ as a function of $t_{\mathrm{Ir}}$. \label{Fig-KerrMicroscopy}}

33 — 1402.5876

\caption{\textbf{\secExpOne}: (\subref{fig:discontinuous_1D_1a}) Predictive mean and 95\% confidence bounds for a GP with SE-ARD {\gpkernel} (\textcolor{blue}{blue solid}), a GP with NN {\gpkernel} (\textcolor{red}{red dotted}) and a log-sigmoid {\newmethodshort} (\textcolor{darkgreen}{green dashed}) on the step function of \eq\eqref{eq:discontinuous}. The discontinuity is captured better by an {\newmethodshort} than by a regular GP with either SE-ARD or NN {\gpkernel}s. (\subref{fig:discontinuous_1D_1b}) The 2D feature space~$\latentsSpace$ discovered by the non-linear mapping~$\mappingNo$ as a function of the input~$\inputsSpace$. The discontinuity of the modeled function is already captured by the non-linear mapping~$\mappingNo$. Hence, the mapping from feature space~$\latentsSpace$ to the output~$\targetsSpace$ is smooth and can be easily managed by the GP.}

34 — 1402.6312

\caption[]{(Color online)(a) Energy \textit{E} = 5 meV cuts at temperatures \textit{T} indicated. Instrumental resolution is shown as a dashed line. Solid lines are the deconvolutions of the constant \textit{E}-cuts into the sum of two Lorentzians. Inset: \textit{T}-dependence of the spatial correlation length $\xi$=$\Gamma^{-1}$ in units of the lattice constant \emph{a}, for \textit{E} = 5 meV (\textbullet), 10 meV ({\color{red}{$\blacktriangle$}}), 15 meV ({\color{green}$\blacksquare$}). (b) S({\bf{q}},\textit{E}) at 5 K for incident energy of 250 meV. Dashed white line emphasizes the spin wave dispersion $\epsilon$({\bf{q}}) connecting the (110) and (210) zone centers. (c) \textit{E}-cuts near the (100) AF zone center summed over the indicated ranges. Solid lines are fits to the sum of two Lorentzians. Inset: Wave vectors of spin waves $\Delta${\bf{q}}, measured relative to (100), for different \textit{E}. Solid line is theoretical expression for $\epsilon$({\bf{q}}) in $\Gamma$-X direction, with \textit{SJ}$_{1}$ = 34 meV. \textit{S} is total spin, \textit{J}$_{1}$ is the nearest neighbor exchange interaction (d) \textit{E}-cuts near (210) averaged on the interval 40-50 meV. Solid lines represent the theoretical lineshape expected for the powder average of the $\epsilon$({\bf{q}}) for the values of J$_{1}$ indicated. (e) Left: Calculations of $\epsilon$({\bf{q}}) along different directions in reciprocal space for values of \textit{J}$_{2}$/\textit{J}$_{1}$ indicated. \textit{J}$_{2}$ is the next-nearest neighbor exchange interaction. Right: Comparison of the experimental density of states DOS (green shaded area) to the powder average of $\epsilon$({\bf{q}}) for values of \textit{J}$_{2}$/\textit{J}$_{1}$ indicated. The low energy part of the DOS is attributed to phonons. (f) Magnetic susceptibility of a collection of single crystals with field applied in ab plane ($\chi_{ab}$) and c direction ($\chi_{c}$) and the powder average $\langle$$\chi$$\rangle$. Orbital susceptibility $\chi_{orb}$ is subtracted from all data. Dashed line shows \textit{T}$_{N}$ = 375 K.}

\caption[]{(Color online) (a) (Log(Transmission)/wavenumber)$^{2}$ for temperatures T = 295 K (\textbf{-}), 325 K ({\color{red}\textbf{-}}), 350 K ({\color{green}\textbf{-}}), 380 K ({\color{cyan}\textbf{-}}), 425 K ({\color{Brown}\textbf{-}}), 450 K ({\color{NavyBlue}\textbf{-}}), 500 K ({\color{Purple}\textbf{-}}). Dashed lines are fits to the 295 K and 500 K data as described in the text. Inset: Raw transmission data for 295 K (black) and 500 K (purple) (b) A comparison of the charge gap $\Delta$ extracted from fits to the optical transmission data, the direct gap in the antiferromagnetic and paramagnetic states determined from DFT+DMFT calculations, and the AF correlation length $\xi$ in units of the in-plane lattice constant \textit{a} extracted from neutron scattering measurements as described in the text.}

35 — 1402.6536

\caption{Comparison between the cumulative distribution functions of normal distributions ({\color{red}red}) and (a) $X$, (b) $\tilde{X}$, (c) $\dot{X}$ when $X\sim \mbox{Bin}(7,1/2)$.}

36 — 1402.6651

\caption{(Color online). Double differential cross section of the process $e+{^{16}}\textrm{O} \to e^\prime + X$ in the quasielastic channel. The calculations have been carried out using Eqs.~\eqref{MC1}--\eqref{MC3} with 20,000 $(p,E)$ pairs sampled from the probability distribution of Eq.~\eqref{prob:dist} and the spectral function of Ref.~\cite{ PhysRevD.72.053005}, without including any modeling of final state interactions. The data, taken from Refs.~\cite{Anghinolfi:1996vm,Anghinolfi:95n} and available online at {\color{blue}{\protect \url{http://faculty.virginia.edu/qes-archive/index.html}}}, are not corrected to remove the effects of FSI.\label{F1}}

\caption{(Color online).~Double differential electron-nucleus cross section in the quasielastic channel. The curves labeled SF have been obtained using Eqs.~\eqref{MC1}--\eqref{MC3} and the model spectral functions of Refs.~\cite{LDA} (for carbon) and~\cite{Ankowski:2007uy} (for calcium and argon). The data are taken from Refs.~\cite{12C2} (for carbon),~\cite{Williamson:1997} (for calcium), and~\cite{Anghinolfi:95n} (for argon). Carbon and calcium data are available online at {\color{blue}{\protect \url{http://faculty.virginia.edu/qes-archive/index.html}}.}\label{F11}}

37 — 1402.6718

\caption{Splashing threshold pressure $P_T$ versus liquid viscosity $\mu$ for silicone oil drops. The impact velocity was fixed at $u_0$=$4.0\pm0.1$ m/s with drop diameter $D$= $1.1\pm0.1$ mm ({\color{red}$\blacktriangle$}), $1.8\pm0.1$ mm ({\color{blue}$\blacksquare$}) or $3.2\pm0.1$ mm ({\color{black}$\bullet$}), fixed at $u_0$ = $2.0\pm0.1$ m/s with D= $3.2\pm0.1$ mm ({\color{black}$\circ$}) or fixed at $u_0$ = $3.0\pm0.1$ m/s with D= $1.8\pm0.1$ mm ({\color{blue}$\Box$}).}

\caption{Thin-sheet ejection time $t_{ejt}$ versus $P$ for $\mu$ = 0.8 ({\color{blue} $\triangle$}), 1.1 ({\color{green}$\lozenge$}), 1.7 ({\color{red}$\circ$}), 2.3 ($\Box$), 4.2 ({\color{blue} $\blacktriangle$}), 5.7 ({\color{green}$\blacklozenge$}), 9.3 ({\color{red}$\bullet$}), and 19.0 ($\blacksquare$) mPa$\thinspace$s silicone oils. The impact velocity and drop size are fixed at $4.0\pm0.1$ m/s and $3.2 \pm 0.1$ mm, respectively. $t_{ejt}(P)$ increases with liquid viscosity. The dashed line serves as a guide to the eye between low-$\mu$ and high-$\mu$ regimes.}

38 — 1402.6974

\caption{A frame showing {\color{red}two translates} of the hyperplane $H$.}

39 — 1402.7193

\caption{Average error for the CZ gate as function of the DC offset error $\Delta\omega_{\text{b},i}$ introduced by miscalibrating the AWG's output to the qubit frequency. Ad-HOC greatly improves the fidelity of the pulse as can be seen by comparing Fig. \textcolor{blue}{a)} and \textcolor{blue}{b)}. The central region of high fidelity does not change since the target fidelity for the calibration protocol was 99.9\%. \label{Fig:Fig4b}}

40 — 1403.0019

\caption{Dependence of the inhomogeneous dephasing time $T_2^*$ on rotation energy $E_{tot}=\sqrt{J^2+h^2}$, where $J$ is the exchange coupling and $h$ is the energy corresponding to the magnetic field difference between the dots. Inset: plot of the extracted values of $T_{2}^{*}$ for $h\simeq$\60.5~neV. Red data points are$T_{2}^{*}$ values obtained using the exchange pulse sequence (Fig.\3), while blue data points are$T_{2}^{*}$ values obtained using the $\Delta B$ pulse (Fig.\2). Main panel:$T_{2}^{*}$ plotted vs.\$E_{tot}$ for $h\simeq 32$~neV, extracted from data shown in SI Appendix Fig.~S3. Red data points are obtained using the exchange pulse sequence (Fig.\S3(b)), and blue data points are obtained using the$\Delta B$ pulse (Fig.\S3(a)). The solid lines in the main panel and in the inset are plots of Eq.\(2) with the same values of $\delta \varepsilon$, the rms fluctuation in the detuning, and $\delta h$, the rms fluctuation of the magnetic field difference, which were obtained by fitting the data for $T_{2}^{*}$ as function of $E_{tot}$ at $h\simeq$~32neV to Eq.\(2). The good agreement of the same form with both data sets is strong evidence that the inhomogeneous dephasing is dominated by charge noise and hyperfine fields and does not depend on the magnetization of the micromagnet. }

41 — 1403.0711

\caption{Weak-scaling property of the method is shown. Elapsed time $T$ is plotted using generated $M=10^4$ samples as a function of the number of cores for various unit problem size $K=64,128,256,512,1024,2048,4096,8192$ and $16384$. For larger $K$, the plotted lines are almost flat. For smaller $K$, slightly increase due to use of the collective communication {\tt mpi\_reduce}. \label{weak}}

42 — 1403.1631

\caption{Top 8 most discriminative events for different stages of exploit execution (Each event set consists of 4 event names in {\color{RoyalBlue}\textsc{Bold}}. E.g, monitoring event set \textit{A-0} consists of simultaneously monitoring \textsc{Ret}, \textsc{Call\_D}, \textsc{Store} and \textsc{Arith} event counts.)}

43 — 1403.1709

\caption{\label{Fig11}Standard uncertainty on the mean value of the sample distributions versus data taking time, for the ``muon target" in position B1 (\opencircle), B2 (\fullsquare) and B3 (\opentriangle).}

44 — 1403.2251

\caption{Porous matrices for fabrication of optical nanowire arrays (a) circular, (b) triangular and (c) square cross-sections [\textcolor{blue}{15}]-[\textcolor{blue}{17}]}

\caption{Transmission Coefficient (TM Modes) patterns for different radius of Ag rod at an operating wavelength of 614THz. In each case the length of each rod and the periodicity are kept same as the Ref. [\textcolor{blue}{12, 13}] while the radius varied from 6nm to 14nm with 1nm resolution (a-i).}

45 — 1403.3067

\caption{\label{fig:lh3}(color online) Ground-state energy of s-shell hypernuclei obtained with the LO chiral YN interaction with cutoff $600\,\text{MeV}/c$. Solid symbols represent J-NCSM results, crosses show IT-NCSM results. Panel (a) shows the ground-state energies of $\isotope[3][\Lambda]{H}$ for $\hbar\Omega=20\,\text{MeV}$, $\alphaYN=0\,\text{fm}^4$ and $\alphaNN=0\,\text{fm}^4$(\squaremark[fill]) and $\alphaNN=0.08\,\text{fm}^4$ (\circlemark[fill]) with EFT-motivated extrapolations (colored bands) compared to the experimental value (gray band) and the result of a Faddeev calculation \cite{Haidenbauer2007} (\LOsixdash, see inset). Panels (b) and (c) show results for the $0^+$ ground states (\squaremark[fill]) and $1^+$ first excited states (\circlemark[fill]) of $\isotope[4][\Lambda]{H}$ and $\isotope[4][\Lambda]{He}$, respectively, using $\alphaYN=\alphaNN=0\,\text{fm}^4$ and $\hbar\Omega=28\,\text{MeV}$. The upper plots show absolute energies, the lower plots excitation energies. The colored bands give the result of an exponential extrapolation of the ground-state energy and the solid lines represent results of previous few-body calculations \cite{Haidenbauer2007}.}

\caption{\label{fig:lli7}(color online) Absolute and excitation energies of the first four states of \isotope[7][\Lambda]{Li} for the LO chiral (b) and the J\"ulich'04 YN interaction (c) compared to the non-strange parent nucleus\isotope[6]{Li} (a). For the LO chiral YN interaction in panel (b) we use the two cutoff values $600\,\text{MeV}/c$ (\LOsixdash) and $700\,\text{MeV}/c$ (\LOsevendash). Experimental data from Refs.~\cite{Davis2005,Hashimoto2006,NuDat2}. All calculations use $\alphaNN=0.08\,\text{fm}^4$, $\alphaYN=0.0\,\text{fm}^4$, and $\hbar\Omega=20\,\text{MeV}$. }

46 — 1403.3145

\caption{(a) A 1.7 mPa$\thinspace$s silicone oil drop is shown impacting smooth glass at 101 kPa (left) and near $P_T$ at 30 kPa (right). Splashing is suppressed at low pressures. (b) \textit{Inset}: $P_T$ vs. $u_0$ of silicone oil drops (R = $1.6\pm0.1$ mm, $\mu_L$ = 1.7 mPa$\thinspace$s) in an atmosphere of air ({\color{red}$\blacktriangle$}) or $SF_6$ ($\triangledown$). The error bars indicate the pressure range for which the ejected sheet first breaks up into droplets. $u^*_0$ indicates the transition between low-$u_0$ and high-$u_0$ regimes. Higher $m_G$ ($SF_6$) lowers the curve but does not affect the trend in data. \textit{Main}: Scaled threshold pressure, $P_T(m_G/m_{air})^{0.5}$, versus $u_0$ for the two gases, collapsing data in both regimes.}

47 — 1403.3153

\caption{Doubly differential cross-sections for electron emission in collisions of 440~keV/u Li$^{q+}$ (q=1,2,3) on He targets as a function of the electron energy for three different emission angles: a) $\theta = 0^\circ$, b) $\theta = 40^\circ$ and c) $\theta = 120^\circ$. \fullsquare: Li$^{+}$; \textcolor{red}{\opencircle}, Li$^{2+}$ ; \textcolor{green}{\opentriangle}, Li$^{3+}$ \cite{Monti2012JPBp5202}. \label{liqmheexp} }

\caption{Doubly differential cross-sections for electron emission in collisions of 440~keV/u Li$^{+}$ on He targets as a function of the electron energy for three different emission angles: a) $\theta = 10^\circ$, b) $\theta = 52^\circ$ and c) $\theta = 150^\circ$. \fullcircle: Experiment; \textcolor{blue}{\dotted}, eTI; \textcolor{green}{\chain}, ePI; \textcolor{magenta}{\dashed}, PTI; \textcolor{orange}{\full}, Total emission. \label{li1mheth}}

\caption{Doubly differential cross-section for electron emission in collisions of 440~keV/u Li$^{2+}$ on He targets as a function of the electron energy for three different emission angles: a) $\theta = 10^\circ$, b) $\theta = 52^\circ$ and c) $\theta = 150^\circ$. % Left panel, CTMC results; Center panel, CDW results, and right panel, CDW-EIS calculations. \fullcircle: Experiment; \textcolor{blue}{\dotted}; eTI; \textcolor{green}{\chain}: ePI; \textcolor{magenta}{\dashed}: PTI; \textcolor{orange}{\full}: Total emission. \label{li2mheth}}

48 — 1403.4049

\caption{{\bf AccMix cell for zinc chloride.} The reactions taking place in the two steps of the AccMix cycle are shown. Legend: \colorbox{blue}{\color{blue}{O}} Concentrated zinc chloride solution; \colorbox{cyan}{\color{cyan}{O}} Dilute zinc chloride solution; \colorbox{red}{\color{red}{O}} Zinc electrode; \colorbox{cyan}{\color{magenta}{X}} Porous silver chloride; \colorbox{green}{\color{green}{O}} Silver. \label{fig:double:CAPMIX}}

49 — 1403.4294

\caption{\label{skeleton} Every point in this skeleton grid has a unique value for branching ratios ${\mathcal B}_A$, ${\mathcal B}_B$, and ${\mathcal B}_C$. Each particular branching ratio decreases from 1 at a vertex to 0 at the side opposite to that vertex. The grid lines are drawn to show the variation of the branching ratios in each direction inside the triangle. The point marked by a {\color{blue} $\bigstar$} denotes the centroid of the triangle and is composed of equal branching ratios, $({\mathcal B}_A, {\mathcal B}_B, {\mathcal B}_C) = (33\%,33\%,33\%)$. The point marked by {\color{red} $\bullet$} is composed of branching ratios, $({\mathcal B}_A, {\mathcal B}_B, {\mathcal B}_C) = (60\%, 20\%, 20\%)$. }

50 — 1403.5274

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Top panel:} the probability distribution function (PDF) of the amplitude of the time derivative of specific angular momentum of the central $3$kpc radius stellar regions (see Eq ~\ref{eq:djdt}) for galaxies of total stellar mass in the range $10^{11}-10^{12}\msun$ in three different redshift ranges, $z=0.62-1$ (black histograms), $z=1-2$ (red histograms), $z=2-3$ (green histograms), respectively. As an intuitive example, if a Milky Way-like galaxy of size $10$kpc and rotation velocity of $200$km/s changes its spin direction by 90 degress in one current Hubble time, it would correspond to a value $\log|dj/dt|$ equal to $2.3$ in the x-axis. {\color{burntorange}\bf Middle panel:} same as the top panel but for the central $7$kpc radius stellar region. {\color{burntorange}\bf Bottom panel:} same as the top panel but for gas in the central $3$kpc radius region. }

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Top panel:} shows the cumulative PDF (CPDF) of the time taken to change the direction of spin of the central $3$kpc radius stellar region by $1$ degree of arc, $t_1$ (Eq ~\ref{eq:t1}) for galaxies of total stellar mass in the range $10^{10}-10^{11}\msun$ (blue curves) and $10^{11}-10^{12}\msun$ (red curves) in three different redshift ranges, $z=0.62-1$ (solid curves), $z=1-2$ (dotted curves), $z=2-3$ (dashed curves), respectively. {\color{burntorange}\bf Middle panel:} shows the of $t_1$ for galaxies of total stellar mass in the range $10^{10}-10^{11}\msun$ in low-density ($\delta_{0.5}=1-10$); solid curves) and high-density environment ($\delta_{0.5}=10^2-10^3$); dotted curves), in two redshift rangess, $z=0.62-1$ (black curves) and $z=1-2$ (magenta curves), respectively. The environment overdensity $\delta_{0.5}$ is defined to be the overdensity of total matter in a sphere of radius $0.5h^{-1}$Mpc comoving. {\color{burntorange}\bf Bottom panel:} shows the CPDF of $t_1$ for blue ($g-r<0.6$; blue curves) and red ($g-r>0.6$, red curves) galaxies of total stellar mass in the range $10^{10}-10^{11}\msun$ in three different redshift ranges, $z=0.62-1$ (solid curves), $z=1-2$ (dotted curves), $z=2-3$ (dashed curves), respectively. }

51 — 1403.5372

\caption{The \gray{} energy spectra multiplied by $E^{2.75}$. The background-sky spectrum (triangles) has been subtracted from the limb spectrum (circles). The scaled local H\,{\sc i} emission\textsuperscript{\ref{Hi}} (crosses) is shown for comparison. Best-fit \gray{} results from two CR models based on the \emph{K\&O} model \cite{KO2012} are also plotted as dotted and dashed lines. Statistical and total (quadrature sum of statistical and systematic) errors are shown as bars and bands, respectively.}

\caption{Best-fit single power law or SPL (top), and best-fit broken power law or BPL (bottom) spectrum for the local CR proton spectrum (solid red lines) as derived from the Earth's limb \gray{} data using the \emph{K\&O} model \cite{KO2012} for $pp$-interactions. The total (combined statistical and systematic, neglecting errors in absolute normalization) uncertainties are the dashed red lines. Other direct measurements (\cite{Panov2009},\cite{CREAM2011pHe},\cite{PAMELA2011},\cite{AMS2000},\cite{BESS2004}) are shown for comparison. The gray band is PAMELA's total uncertainty.}

52 — 1403.5418

\caption{(Colour online). Cartoon picturing the dynamics of the rings. The rings in the gel (left) are represented by lattice animals moving on the dual of the gel lattice (centre). \textbf{(a) (b)} and \textbf{(c)} Represent the moves allowed in a gel: translation, bending and branching, each by retraction of a terminal segment (shaded in grey). See text for more details. \textbf{(d)} With probability $p_{th} p_{m \rightarrow m^\prime}^{branch/bend}$ a bead can move onto a site that is already occupied to become an effective \emph{pin} (green) for the occupant bead $m_{p}$ (red). \magenta{Computationally, we implement this by stacking two beads on top of each other, on the same site (green bead on top of red one)}. \textbf{(e)} This is the case in which the threaded segment coincides with an end, in this case such end becomes ``pinned''. \magenta{In this configuration, the red segment/bead cannot be retracted, while the green one can contribute to the motion by annihilating with an anti-kink (see text). In this case $m^\prime$ can extend only if one of the grey beads or the green one is removed. In the case the red one is attempted to be retracted, the move would be rejected.} \textbf{(f)} and \textbf{(g)} show two snapshots of Molecular Dynamics simulation showing a self-threading and a self-trapping configurations, respectively. The colors highlight different segments of the chain. In (f) an ending segment of a branch (green) threads through another ending segment of another branch (red). This case is analogous to case (e) in the left panel. Fig. (g) is obtained after a strong electric field is applied to a self-threading configuration. One can imagine applying a field directed upward to the configuration in (d). The green segment may elongate, while the grey (free) end slides backward until it coincides with the red segment. If this occurs, the configuration is ``trapped'', like the one pictured in (g). The gel structure in (f) and (g) is sketched only partially and thinned for simplicity. \magenta{(see S.I. for details)}}

\caption{(Colour online). Mean square displacement of the centre of mass of the rings as a function of the time for different values of the threading probability $p_{th}$ and increasing length $M$. We observe sub-diffusive behaviour $\delta^2 r_{CM} \sim t^{x}$ with $x<1$ \blue{until intermediate times}, even at moderate probability of threading $p_{th}$. The slowing down can only be caused by an increasing number of self-threadings, which become more important for longer rings (see text for details). In clockwise order: (a) $M=32$, (b) $M=64$, (c) $M=128$ and (d) $M=256$ beads. \blue{(see text for details)}.}

\caption{(color online). Fraction of threadings present $t$ time-steps after that $p_{th}$ is turned off for $M=512$ and different values of $p_{th}$ used to bring the system to equilibrium. $G_{th}(t)$ shows a slow decay that lasts for several decades. This suggests the presence of a hierarchical structure of self-threadings and long-lived correlations affecting the long time dynamics. \red{For the longest rings all the threadings did not disappear in the simulation runtimes accessible to us.} (Inset) $G_{th}(t)$ for $p_{th}=1$ and different values of $M$.}

\caption{(color online). Numeric integral of $G_{th}(t)$, $\eta_{th} = \int_0^{\infty} G(t) dt$. Independently on the value of $p_{th}$, the data points are fitted by the power law $\tau_{th} \sim M^\alpha$, with $\alpha = 3$ for $M<128$ and $\alpha=3.5$ for $M>128$. \red{The arrow at M=512 indicates that the values of $\eta_{Th}$ for the longest rings represent only a lower bound as we could not observe the removal of all the threadings within the simulation runtime.} }

53 — 1403.6167

\caption{Cable system of Sec.~\ref{sec:Case1}: positive-sequence resistance (top panel) and inductance (bottom panel) computed using FEM ({\color{blue} ${\bf \mathlarger \circ}$}), MoM-SO ($\oldcdot[-1.5pt]$), and cable constant ({\color{red} -\,-}). Screens are continuously grounded.}

\caption{Cable system of Sec.~\ref{sec:Case1}: positive-sequence resistance (top panel) and inductance (bottom panel) computed using FEM ({\color{blue} ${\bf \mathlarger \circ}$}), MoM-SO ($\oldcdot[-1.5pt]$), cable constant with Pollaczek ground return ({\color{red} \xdash[0.5em] \xdash[0.5em]}), cable constant with Saad ground return ({\color{red} \xdash[0.1em] \xdash[0.1em] \xdash[0.1em] \xdash[0.1em]}) and MoM-SO with approximate ground return effects~\cite{TPWRD2} ({\color{magenta} \xdash[1.2em]}). The screens of the cables are open.}

\caption{Cable system considered in Sec.~\ref{sec:results_case1c}: resistance (top panel) and inductance (bottom panel) computed using FEM ({\color{blue} ${\bf \mathlarger \circ}$}), MoM-SO ($\oldcdot[-1.5pt]$), cable constant with Pollaczek ground return ({\color{red} \xdash[0.5em] \xdash[0.5em]}), cable constant with Saad ground return ({\color{red} \xdash[0.1em] \xdash[0.1em] \xdash[0.1em] \xdash[0.1em]}), and MoM-SO with approximate ground return effects \cite{TPWRD2} ({\color{magenta} \xdash[1.2em]}). Phase conductors are open, and current is injected in the sheaths.}

\caption{System of three SC cables in a tunnel considered in Sec.~\ref{sec:results_case2}: resistance (top panel), and inductance (bottom panel) computed with FEM ({\color{blue} ${\bf \mathlarger \circ}$}) and MoM-SO ($\oldcdot[-1.5pt]$). In order to show the effect of the tunnel, the resistance and inductance of the cables buried directly in ground are also shown ({\color{red} $\times$}).}

54 — 1403.6633

\caption{Entries are: $N$, the number of KBOs; $m$, the mass of the single KBO, in units of Earth mass; $R$, the radius of the single KBO; $T$, the time-scale to repopulate the \textcolor{blue}{low-eccentricity} KBO distribution. Actually, for the simulations $\#$ 5-6-7 we checked only the initial cooling phase and then the actual values of the parameter $T$ were extrapolated using Eq.$\,$\ref{eq:rel}. }

55 — 1403.6634

\caption{ Red line with circle (\textcolor{red}{$\multimap$}) shows the regime of Hydrodynamic turbulence with its starting point at $R_e \approx 2500$ \cite{Avila2011}. Black dots ($\bullet$) show the regime of elastic turbulence simulation by Berti et. al. using Oldroyd-B model \cite{berti_10}. Blue circle, square and pentagon (\textcolor{blue}{$\circ,\Box, \bigstar$}) are the points on $R_e-W_i$ plot where authors have contributed ; the work has been presented here as well as in some previous publications \cite{sanat_kh_str,sanat_kh_aip13}. Plus ($+$) is point where Molecular dynamics simulations have been performed for strongly coupled Yukawa fluids \cite{Ashwin2010}. Red square (\textcolor{red}{$\Box$} point outs regime where some recent experiments have been performed by Groisman et. al \cite{Groisman_nature}). The shaded square region covers area in which studies made by Burghelea et. al \cite{burghelea_07}.}

56 — 1403.6647

\caption{\textcolor{red}{\large \label{fig:hoe}}(color online) Higher order entanglement is observed using Hillery-Zubairy criteria. Solid line shows spatial variation of $E_{ab_{1}}^{2,1}$ and dashed line shows spatial variation of $E_{ab_{1}}^{\prime2,1}$ with the initial state $|\alpha\rangle|\beta\rangle|\gamma\rangle$ and $k=0.1,\,\Gamma=0.001,\,\Delta k=10^{-4},\,\alpha=5,\beta=2,\gamma=1.$ }

57 — 1403.6886

\caption{Investigating computational issues with pMCMC for the Lotka--Volterra model \red{defined in section~\ref{sec:lvintro}}. (a) The true underlying synthetic data set. Species are observed at discrete time points and corrupted with $N(0, 10^2)$ noise. (b) Twelve trace plots of $\theta_{1}$ from pMCMC chains initialised with random draws from the prior (see expression \ref{eq:prior}). The chains fail to explore the space. (c) shows the median and 95\% interval for estimates of the log--likelihood from the particle filter for varying $\theta_{1}$ close to the true value, for $\theta_{2}$ and $\theta_{3}$ fixed at the true values. (d) shows that the variance og log--likelihood estimates increases away from the true values.}

58 — 1403.6953

\caption{$\Eapp$ is the outer ellipse. The identification ellipse obtained by MOOSE is shown in red ('\textcolor{red}{- -}') and the one obtained by the proposed method in black ('\textcolor{black}{--}').}

59 — 1403.7049

\caption{Spectrum of the S$_{1/2}\rightarrow$ P$_{1/2}$ transition in calcium. The spectrum obtained by scanning the spectroscopy laser without additional cooling intervals ({\color{red}-}${\color{red}\Box}${\color{red}-}) shows a sharp drop close to the line centre. Employing additional cooling intervals results in a full, undistorted spectral line ({\color{blue}-}${\color{blue}\bigcirc}${\color{blue}-}). A Voigt profile is fit to the undistorted spectrum (blue line).}

60 — 1403.7397

\caption[]{Helium abundances for hot HB stars in M\,80, NGC\,5986, NGC\,6752, and$\omega$\,Cen (Moni Bidin et al.), M\,3, M\,13, and NGC\,6752 (Moehler et al.) and NGC\,288. The grey dots mark the abundances predicted for various HB ages from diffusion theory (with an ad hoc surface-mixing zone,\citealt{miri11}, see Sect.\,\ref{ssec:diff} for details). The tracks have dots every 3 Myr.}

\caption[]{\color{black}Abundances derived via spectrum synthesis for all stars hotter than 9\,000\,K. The triangles mark the same stars as in Fig.\,\ref{fig:cmd_hb}. The asterisks mark the results from \citet{behr03}. The bars at 17\,500\,K mark the average error bars of\citet{behr03} for NGC\,288. Our errors are probably not smaller. The dashed lines mark the solar abundances. The four-pointed stars mark the abundances derived for stratified model spectra (left column, see Sect.\,\ref{ssec:strat} for details), which indicate the equilibrium abundances achievable by diffusion. The grey dots (right column) mark the abundances predicted for various HB ages from diffusion theory (with an ad hoc surface-mixing zone, \citealt{miri11}, see Sect.\,\ref{ssec:diff} for details). The tracks have dots every 1 Myr.}

\caption[]{Normalised spectra for stars 199 (upper panel, 12\,400\,K) and 221 (lower panel, 13\,400\,K) compared with stratified model spectra F (upper panel) and H (lower panel), which have fitted parameters close to those of the observed spectra. For clarity, the stratified model spectra have been offset by 0.1 from the observed spectra (see Tables\,\ref{tab:par_spec_hot} and \ref{tab:strat_homo}).\label{Fig:obs_strat}}

61 — 1403.7475

\caption{\label{Fig6} In all panels, unless specified, $L=200$, $\mu=0$, $W=t$ and $\gamma_e=\gamma_{ep}=t$. (a) Thermopower distributions in the VRH regime, when $\mu$ lies in the bulk (left red curve, $V_g=0$, $W=4t$) or close to the edge (right blue curve, $V_g=2.2t$, $W=t$) of the impurity band. Data are given for $L=200$ (full lines) and $L=400$ (circles). The straight dashed lines underline the exponential behaviour of the tails $\sim \text{exp}\{-cS\}$ predicted in Ref.~\cite{Jiang2013}. In both cases, $k_BT=t$. (b) Main panel: Typical thermopower as a function of $T$ around the (lower) band edge. From the bottom to the top, the various curves correspond to $V_g/t=1.5\,\text{({\large{$\circ$}})},1.9\,\text{({\scriptsize{\color{red}$\square$}})}, 2.0\,\text{({\large{\color{ForestGreen}$\diamond$}})},2.1\,\text{({\scriptsize{\color{blue}$\triangle$}})}, 2.2\,\text{({\scriptsize{\color{magenta}$\triangledown$}})}$ and $2.3\,\text{({\scriptsize{\color{brown}$\times$}})}$. Dotted lines are guides to the eye. Inset: zoom at very large temperatures $k_BT\gtrsim E_B$. The fits $f(V_g)/T$ (dashed lines) confirm the expected behaviour $S_0\sim T^{-1}$ (Eq.~\eqref{eq:S_largeT}). (c) Typical thermopower as a function of $V_g$, for $k_BT/t=0.1\,\text{({\large{$\circ$}})},0.2\,\text{({\scriptsize{\color{red}$\square$}})}, 0.5\,\text{({\large{\color{ForestGreen}$\diamond$}})},1.0\,\text{({\scriptsize{\color{blue}$\triangle$}})}, 2.5\,\text{({\scriptsize{\color{magenta}$\triangledown$}})}$ and $10.0\,\text{({\scriptsize{\color{brown}$\times$}})}$. At large $V_g$ (when $\mu$ lies outside the band), dashed lines are linear fits with slope $t/k_BT$ (Eq.~\eqref{eq:S_zvyagin_32}). Dotted lines are guides to the eye. (d) Typical thermopower as a function of $T$, for electron-phonon coupling strength $\gamma_{ep}/t=1$ (full line), $0.5$ (dashed line), $0.1$ (dotted line) and $0.05$ (mixed line), at $V_g=1.9t$ (black curves, bottom set) and $V_g=2.1t$ (red curves, top set). }

62 — 1403.7708

\caption{ (a) Plot of $T_{\rm coll}$ versus $\langle x\rangle$ obtained in the bounded phase, for microtubules. Black squares ($\Box$) are for $N=1$, $f=0$, $c<c_{crit}$, and red circles (\textcolor{red}{$\bigcirc$}) are for $N=1$, $f>f_s^{(1)}$, $c=100 \mu$M$\gg c_{crit}$ (for microtubule parameters, $c_{crit}=8.67 \mu$M). While, for $N=2$, $f>f_s^{(2)}$, values of $T_{\rm coll}$ are obtained at four different concentrations (all greater than $c_{crit}$) $c=9 \mu$M (\textcolor{Brown}{$\blacktriangledown$}), $40 \mu$M (\textcolor{Magenta}{$\bullet$}), $70 \mu$M (\textcolor{Green}{$\blacktriangle$}), and $100 \mu$M (\textcolor{Blue}{$\blacksquare$}). (b) A comparison of three time-traces of the wall-position $x(t)$ in the bounded phase, for parameters: (i) $N=1$, $f=0$, $c=7.5 \mu$M$<c_{crit}$ (black curve); (ii) $N=1$, $f=17.6$ pN$>f_s^{(1)}$, $c=100 \mu$M (red curve); and (iii) $N=2$, $f=35.3$ pN$>f_s^{(2)}$, $c=100 \mu$M (blue curve). Note that $\langle x\rangle$ is nearly same for all trajectories. }

63 — 1404.0194

\caption{(color online) The $J^{XXZ}_{1}$--$J^{XXZ}_{2}$ model on the honeycomb lattice, showing (a) the bonds ($J_{1} \equiv$ ----- ; $J_{2} \equiv \textcolor{blue}{- - -}$) and the two sites (\textcolor{red}{{\Large $\bullet$}}) A and B of the unit cell; (b) the N\'{e}el planar, N(p), state; (c) the N\'{e}el $z$-aligned, N($z$), state; and (d) the N\'{e}el-II planar, N-II(p), state. The arrows represent the directions of the spins located on lattice sites \textbullet.}

64 — 1404.0337

\caption{{\ListRecolor$(G,L,\alpha,\beta,\ell)$}}

65 — 1404.1066

\caption{ Comparison of test error, training time, and speedup of kernelized SVM training methods. The first column indicates dataset file size, number of instances, dimensionality, and SVM hyperparameters $C$ and $\gamma$ (with a citation for previously published values, otherwise derived by cross-validation using GTSVM). Results for SP-SVM are the average of five runs with different randomly sampled candidate sets (see text for standard deviations). Row background colors indicate implementation architecture: single-core (SC), \colorbox{blue1}{multi-core (MC)}, \colorbox{blue2}{GPU}. \textcolor{red1}{Red font color} indicates poor test error results. \textbf{Bold typeface} indicates the best timing results for each dataset and architecture. Symbol $^\dagger$ indicates accuracy metric is $(1-$AUC$)\%$. % with false positive rate less than $10\%$. Symbol \na{} indicates a data set/method pair that was unable to be run, as explained in the text. }

66 — 1404.2068

\caption{Distributions of $\rho_{ch}$s with respect to distance from the shower core of $\gamma$, proton and iron primaries at different energies given by the QGSJET-GHEISHA model combination over the observation levels of Hanle (black symbols) and Pachmarhi (red symbols). In the respective plots, {\large $\bullet$/\textcolor{r}{$\bullet$}} indicates for 100 GeV $\gamma$, 250 GeV proton and 5 TeV iron primaries; $\blacktriangle$/\textcolor{r}{$\blacktriangle$} indicates for 500 GeV $\gamma$, 1 TeV proton and 10 TeV iron primaries; and $\bigstar$/\textcolor{r}{$\bigstar$} indicates for 1 TeV $\gamma$ and 2 TeV proton primaries.}

\caption{Distributions of offset normalised $t_{ch}$ with respect to distance from the shower core of $\gamma$, proton and iron primaries at different energies given by the QGSJET-GHEISHA model combination over the observation levels of Hanle (black symbols) and Pachmarhi (red symbols). In the respective plots, {\large $\bullet$/\textcolor{r}{$\bullet$}} indicates for 100 GeV $\gamma$, 250 GeV proton and 5 TeV iron primaries; $\blacktriangle$/\textcolor{r}{$\blacktriangle$} indicates for 500 GeV $\gamma$, 1 TeV proton and 10 TeV iron primaries; and $\bigstar$/\textcolor{r}{$\bigstar$} indicates for 1 TeV $\gamma$ and 2 TeV proton primaries.}

67 — 1404.2193

\caption{The SEDs (top) and light curves (bottom) in the pure Synchrotron Self-Compton (SSC) case with change of the magnetic field as the cause of the flare. {\bf Top:} The colored triangles are simultaneous data points from SMARTS, \swift and \fermi, with blue/red being the pre-flare/flaring states. The grey squares and circles are historical data. The open blue/red triangles are part of flare 1 identified by \citet{chatterjee_2013:0208_opticalonly}, while the filled blue/red triangles are for flare 2. The \swift bow-tie and the \fermi upper limits are simultaneous with the filled red triangles. % The grey data are not strictly simultaneous with the colored data. The three histograms show the simulated SEDs before, during, and after the peak of flare 2, and so they should match the filled triangles, while other data points are plotted there for reference. The dotted magenta line is the steady thermal emission component. The dashed orange red lines refer to the first second order SSC emission during the peak of the flare. {\bf Bottom:} In the lower panel the open circles show the 10-day-averaged optical light curves in B (blue) and J (red) bands during flare 2. The histograms show two simulated synchrotron light curves at similar frequencies. In the upper panel there are no observational data points. Fermi \gray flux was always below detection in 10-day bins during this time. Three simulated IC light curves are shown, with green, orange, and black solid lines representing the energy bands in \xray, \fermi \gray, and very high energy (VHE) \gray. The shaded grey areas mark the phase when the simulation is still in setup phase. The vertical dotted line marks the peak of the synchrotron flare.}

68 — 1404.2380

\caption{The three network regions considered. \textcolor{NewText}{The location of the reference receiver is indicated by ($\ast$).} \label{Figure:Arenas} }

69 — 1404.2474

\caption{Time history of energy and angular momentum. The total energy (blue) of the flux tube is conserved. Part of the tube inside the ergosphere loses energy, i.e.\gains negative energy (black), while other parts of the tube compensate this loss with positive energy (red). Although the flux tube continuously falls into the ergosphere, after some time$\approx50\,r_h/c$ with $r_h$ the radius of the event horizon, the energy content of the flux tube outside the ergosphere (green) exceeds the total initial energy of the flux tube (blue). In the lower panel the corresponding angular momentum evolution is shown. }

70 — 1404.2887

\caption{{\bf Evolution of unstable eigenfunctions.} (a,d) show $\|\bu(t) \|$ versus time $t$ for small perturbations of the doubly-localized solution along its most unstable eigenfunctions, at (a) $\Rey = 380$ and (d) $\Rey = 400$. Color denotes the symmetries of the eigenfunction perturbations: {\color{red} red} for \revision{$\langle \sxy, \sz \rangle$} symmetric eigenfunctions, black for \revision{$\langle \sxy, -\sz \rangle$}, {\color{blue} blue} for \revision{$\langle -\sxy, \sz \rangle$}, {\color{green} green} for \revision{$\langle -\sxy, -\sz \rangle$}, and {\color{cyan} cyan} for perturbations along combinations of eigenfunctions that break all symmetries. Solid lines indicate the most unstable eigenfunction of each symmetry group, and dashed are the second-most unstable. The midplane streamwise velocity of the most unstable perturbation at (b) $t=0$ and (c) $t=200$ for $\Rey=380$, and (e,f) the same for $\Rey=400$. }

71 — 1404.3184

\caption{Illustration of computing $\Omega^*_{\grave{\bf w}} \left({\bf x}\right)$ in the 2-dimensional case.} \label{fig:figure2} \end{figure} Since it is clear that $\Omega_{\grave{\bf w}} ({\bf u})$ depends only on the magnitudes of the components of ${\bf u}$ we may limit our attention to the first orthant. Moreover, since $\Omega_{\grave{\bf w}} ({\bf u})$ is invariant to permutations of the components of ${\bf u}$ (because of the underlying sorting operation), we may consider, without loss of generality, some vertex ${\bf c} = [c_1,c_2,...,c_n]$, satisfying \begin{equation} c_1 \geq c_2 \geq \cdots \geq c_n \geq 0.\label{eq:sorted_entries} \end{equation} The vertices of a convex polyhedron are those points in the polyhedron that can not be obtained as convex combinations of other points; that is, such that the set is the convex hull of that set of points. In the case of ${\cal S}$, the subset of vertices in the first orthant, with components satisfying \eqref{eq:sorted_entries}, is \begin{equation} \tilde{\cal C} = \{{\bf c}_1, {\bf c}_2, \cdots, {\bf c}_n\}, \end{equation} with \begin{equation} \begin{split} {\bf c}_1 &= [\tau_1, 0, 0, \cdots, 0 ]^T \\ {\bf c}_2 &= [\tau_2, \tau_2, 0, \cdots, 0 ]^T \\ \vdots\\ {\bf c}_{n} &= [\tau_{n}, \tau_{n}, \cdots,\tau_{n}, \tau_{n}]^T \\ \end{split} \end{equation} where \begin{equation} \tau_i = \left( \sum_{j=1}^i \grave{w}_j\right)^{-1}\!\!, \;\; i \in \{1, \cdots, n \}. \end{equation} It is clear that these points belong to ${\cal S}$ and that none of them can be obtained as a convex combination of the others. Figure \ref{fig:figure2} depicts the 2-dimensional case. The complete set of vertices ${\cal C}$ (in the general case, see the next paragraph for exceptions) is obtained by considering all possible permutations of the entries of each element of $\tilde{\cal C}$ and all possible sign configurations. Notice that for certain particular choices of $\grave{\bf w}$, some of the points mentioned in the previous paragraph are no longer vertices. For example, if $\grave{w}_1 = ... = \grave{w}_2 = w$, we have $\tau_i = (i w)^{-1}$; in this case, only the points that have exactly one non-zero component are vertices, since in this case we recover the $\ell_1$ norm. Given some vector ${\bf x}\in \mathbb{R}^n$, let ${\bf x}_{(k)}$ denote the vector obtained by keeping the $k$ largest (in magnitude) elements of its sorted version $\grave{\bf x}$ and setting the others to zero. Also, let $x_{(k)}$ be the $k$-largest element of ${\bf x}$ in absolute value. Naturally, $\left\|{\bf x}_{(1)}\right\|_1 = |x_{(1)}| = \left\|{\bf x}\right\|_{\infty} $ and $\left\|{\bf x}_{(n)}\right\|_1 =\left\|{\bf x}\right\|_1$. Therefore, we have \begin{equation} \label{dual_norm} \begin{split} \Omega^*_{\grave{\bf w}} \left({\bf x}\right) &= \max_{{\bf u} \in {\cal S}} \left\langle {\bf u}, {\bf x}\right\rangle\\ & = \max_{{\bf c} \in {\cal C}} \left\langle {\bf c},\grave{\bf x}\right\rangle \\ &= \max \{ \langle {\bf c}_k,|{\bf x}_{(k)}|\rangle, k =1, \cdots, n \}, \end{split} \end{equation} where $|{\bf v}|$ denotes the vector obtained by taking the absolute values of the components of ${\bf v}$. The derivation in \eqref{dual_norm} proves the following theorem. \vspace{0.15cm} \begin{theorem} The dual norm of $\Omega_{\grave{\bf w}}$ is given by \begin{equation} \label{dual_norm_final} \Omega_{\grave{\bf w}}^* ({\bf x}) = \max \Bigl\{\tau_k \bigl\|{\bf x}_{(k)}\bigr\|_1,\;\; k = 1, \cdots, n\Bigr\}. \end{equation} \end{theorem} \vspace{0.15cm} As a simple corollary of this theorem, notice that, in the case where $\grave{w}_1 = ... = \grave{w}_2 = w$ (that is, if $\Omega_{\grave{\bf w}} ({\bf x}) = w\|{\bf x}\|_1$), it results from \eqref{dual_norm_final} that \begin{equation} \label{dual_norm_l1_linf} \Omega_{\grave{\bf w}}^* ({\bf x}) = \max \biggl\{ \frac{\bigl\|{\bf x}_{(k)}\bigr\|_1}{k\, w} ,\;\; k = 1, \cdots, n\biggr\} = \frac{1}{w} \, \| {\bf x}\|_{\infty} \end{equation} because \[ |x_{(1)}| \geq \tfrac{1}{2}(|x_{(1)}|+|x_{(2)}| ) \geq ... \geq \tfrac{1}{n}(|x_{(1)}|+ ... + |x_{(n)}| ), \] recovering the well-known fact that the dual of the $\ell_1$ norm is the $\ell_{\infty}$ norm. %Now, we are ready to give the subdifferential of $\Omega_{\grave{\bf w}}$: %\begin{lemma} The subdifferential of $\Omega_{\grave{\bf w}}$ is given by %\begin{equation}\label{subdif_l1} % \partial \Omega_{\grave{\bf w}}\left({\bf x}\right) = \left\{ %\begin{split} %&{\cal D}, \;\;\; \mbox{if} \; {\bf x} = 0\\ %&\left\{{\bf u} \in \mathbb{R}^n: % \begin{array}{ll} % {\bf u} \; \mbox{at the boundary of} \\ % {\cal D}, \mbox{and}\; {\bf u}^T{\bf x} = \Omega_{\grave{\bf w}}\left({\bf x}\right) % \end{array}\right\}, \;\;\; \mbox{if} \; {\bf x} \neq 0 \\ %\end{split} %\right. %\end{equation} %\end{lemma} %\proof This lemma is directly derived from Proposition 1.2 in \cite{bach2011optimization} and Lemma 3 above. %\endproof %Next, we further %analyze the relationship between DWSL1 and {\it quadratic program} (QP). %Given $\grave{\bf w}$, \eqref{dwsl1} can be rewritten as a QP: %\begin{equation} %\min_{{\bf x}_+, {\bf x}_+ \in \mathbb{R}^n_+ } \frac{1}{2} \left\| {\bf y} - %{\bf A}\left({\bf x}_+ -{\bf x}_-\right) \right\|_2^2 + {\bf w}^T \left({\bf x}_+ + {\bf x}_-\right) %\end{equation} %with \eqref{w_vector}. Or a {\it variable-shift QP}: %\begin{equation} %\min_{\substack{{\bf x}_+, {\bf x}_+, \\ \grave{\bf x}_+ , \grave{\bf x}_- \in \mathbb{R}^n_+} } %\frac{1}{2} \left\| {\bf y} - {\bf A}\left({\bf x}_+ -{\bf x}_-\right) \right\|_2^2 + %\grave{\bf w}^T \left(\grave{\bf x}_+ + \grave{\bf x}_-\right) %\end{equation} %where $\left\|\grave{\bf x}\right\|_1 = \grave{\bf x}_+ + \grave{\bf x}_-$, note that $\grave{\bf x}$ %is a magnitude-sorted vector while $\grave{\bf x}_+, \grave{\bf x}_-$ are always not. %\subsection{Optimality Conditions and Solution Uniqueness} %\begin{lemma}[Optimality Conditions] %A vector ${\bf x}^* \in \mathbb{R}^n$ is a solution of \eqref{dwsl1} if and only if $\forall j = 1, \cdots, n,$ %\begin{equation}\label{opt_cond} % \left\{ %\begin{split} %&\left|{\bf a}_j^T\left({\bf y} - {\bf A} {\bf x}^* \right)\right| \leq w_j^* &\ \ \mbox{if} \ x^*_j = 0\\ %&{\bf a}_j^T\left({\bf y} - {\bf A} {\bf x}^* \right) = w_j^* \, \mbox{sign} (x^*_j) &\ \ \mbox{if} \ x^*_j \neq 0, %\end{split} %\right. %\end{equation} %where ${\bf a}_j$ denotes the $j$-th column of ${\bf A}$, and $x_j^*$ and $w_j^*$ the $j$-th entry of ${\bf x}^*$ and %\begin{equation} \label{w_x_star} %{\bf w}^* = \left({\bf P}\left({\bf x}^*\right)\right)^T \grave{\bf w}, %\end{equation} %respectively. %\end{lemma} %\vspace{0.2cm} %\proof %The first-order optimality condition of \eqref{dwsl1} yields %\begin{equation} \label{first_order_opt} %{\bf A}^T\left({\bf y} - {\bf A}{\bf x}^*\right) \in \partial \left\|{\bf w}^*\odot{\bf x}^*\right\|_1. %\end{equation} %Now let ${\bf s} \in \partial \left\|{\bf w}^*\odot{\bf x}^*\right\|_1$; then, for $j \in \left\{1, \cdots, n\right\}$, %\begin{equation}\label{subdif_l1} % \left\{ %\begin{split} %&\left|s_j\right| \leq w_j^* &\ \ \mbox{if} \ x^*_j = 0\\ %&s_j =w_j^*\, \mbox{sign}(x_j^*) &\ \ \mbox{if} \ x^*_j \neq 0.\\ %\end{split} %\right. %\end{equation} Combining \eqref{first_order_opt} and \eqref{subdif_l1} establishes \eqref{opt_cond}. %\endproof %\vspace{0.2cm} %\begin{lemma}[Solution Uniqueness] %Let ${\bf x}^* \in \mathbb{R}^n$ be a solution of \eqref{dwsl1}, %$J = \left\{j: \left|{\bf a}_j^T\left({\bf y} - {\bf A} {\bf x}^* \right)\right| = w_j \right\}$, %${\bf A}_J = [{\bf a}_j]_{j \in J}$ the submatrix of ${\bf A}$ whose columns are indexed by $J$, and ${\bf w}$ %as defined as above. %If ${\bf A}_J$ is full column rank, then ${\bf x}^*$ is unique, and we have %\begin{equation} %{\bf x}^*_J = \left({\bf A}_J^T{\bf A}_J\right)^{-1}\left({\bf A}_J^T{\bf y} - {\bf w}_J^* \odot {\bf u}\right), {\bf x}^*_{J^c}=\textbf{0} %\end{equation} %where $J^c$ is the complement of $J$, ${\bf u} = \mbox{sign}\left({\bf A}_J^T\left({\bf y} - {\bf A}{\bf x}^*\right)\right)$, %${\bf x}^*_J$ is a subvector of ${\bf x}^*$ indexed by $J$, and ${\bf w}_J^*$ the subvector of ${\bf w}^*$ (obtained by \eqref{w_x_star}) whose entries are indexed by $J$. %\end{lemma} %\vspace{0.2cm} %\proof %This lemma is a generalization of Lemma 2 in \cite{tibshirani2013lasso} by replacing %the regularization parameter by ${\bf w}^*$, where ${\bf w}^*$ is as given by \eqref{w_x_star}. %\endproof % %\subsection{Dual Problem and Sub-optimality Bound} % %An equivalent problem to \eqref{dwsl1} is %\begin{equation} \label{wsl1_constr} %\begin{split} %&\min_{{\bf z}\in \mathbb{R}^m, {\bf x} \in \mathbb{R}^n} \ %\tfrac{1}{2} {\bf z}^T{\bf z} + \left\|{\bf w} \odot {\bf x}\right\|_1 \\ %& \mbox{subject to} \ \ {\bf z} = {\bf A}{\bf x} - {\bf y}, %\end{split} %\end{equation} %where ${\bf w} = \bigl({\bf P}({\bf x})\bigr)^T \grave{\bf w}$. The Lagrangian for this problem is %\begin{equation} %\mathcal{L}\left({\bf x}, {\bf z}, \boldsymbol{\mu}\right) = %\tfrac{1}{2} {\bf z}^T{\bf z} + \left\|{\bf w} \odot {\bf x}\right\|_1 + %\boldsymbol{\mu}^T\left({\bf A}{\bf x} - {\bf y} - {\bf z}\right) %\end{equation} %where $\boldsymbol{\mu}$ is the Lagrangian multiplier (and dual variable). Thus, the dual problem becomes %\begin{multline} \label{dual_problem} %\max_{\boldsymbol{\mu}} \min_{{\bf x}, {\bf z}} \mathcal{L}\left({\bf x}, {\bf z}, \boldsymbol{\mu}\right) = %\max_{\boldsymbol{\mu}} -\boldsymbol{\mu}^T{\bf y} + %\min_{{\bf z}} \left( \tfrac{1}{2} {\bf z}^T{\bf z} - \boldsymbol{\mu}^T{\bf z}\right) \\ %- \underbrace{\max_{\bf x}\left( -\left({\bf A}^T\boldsymbol{\mu}\right)^T{\bf x}- %\left\|{\bf w} \odot {\bf x}\right\|_1 \right)}_{\Omega_{\grave{\bf w}}^* \left(-{\bf A}^T\boldsymbol{\mu}\right)} %\end{multline} %where the $\min_{\bf z}$ term attains its extremum $-\tfrac{1}{2}\boldsymbol{\mu}^T\boldsymbol{\mu}$ %at %\begin{equation} \label{z_mu} %\hat{\bf z} =\boldsymbol{\mu} %\end{equation} %and the $\max_{\bf x}$ term is the conjugate of $\Omega_{{\grave{\bf w}}} \left({\bf x}\right)$ computed at %$-{\bf A}^T{\boldsymbol{\mu}}$; let its optimum be $\hat{\bf x}$, and we have %${\bf A}^T{\boldsymbol{\mu}} \in {\bf w} \odot \partial \left\|\hat{\bf x}\right\|_1 %%= \left({\bf P}\left(\hat{\bf x}\right)\right)^T \grave{\bf w} \odot % %\partial \left\|\hat{\bf x}\right\|_1 %$ %thus yielding %\begin{equation} \label{dual_norm} %\Omega_{{\grave{\bf w}}}^* \left(-{\bf A}^T\boldsymbol{\mu}\right) = \left\{ %\begin{split} %&0 \ \ \ \mbox{if} \ {\boldsymbol{\mu}} \in \mathcal{C} \\ %&\infty \ \ \ \mbox{otherwise} %\end{split} %\right. %\end{equation} %where %\begin{equation} \label{constraint_set} %\mathcal{C} = \left\{\boldsymbol{\mu}: \; \left|\left({\bf A}^T\boldsymbol{\mu}\right)_i\right| %\leq w_i, i=1, \cdots n \right\} %\end{equation} %is the dual feasibility set. %Combining \eqref{dual_problem} and \eqref{dual_norm} yields the dual problem of \eqref{dwsl1}: %\begin{equation} \label{dual_wsl1} %\max_{\boldsymbol{\mu} \in \mathbb{R}^m} \ %\mathcal{D}\left(\boldsymbol{\mu}\right) = -\tfrac{1}{2}\boldsymbol{\mu}^T\boldsymbol{\mu} - %\boldsymbol{\mu}^T{\bf y} \ \ \mbox{subject to} \ {\boldsymbol{\mu}} \in \mathcal{C}. %\end{equation} %Note that a dual feasible point $\boldsymbol{\mu}$ gives a lower bound on the optimal value ${\bf p}^*$ %of the primal problem \eqref{dwsl1}, {\it i.e.}, $\mathcal{D}\left(\boldsymbol{\mu}\right) \leq {\bf p}^*$, %which is called {\it weak duality}. Furthermore, {\it strong duality} holds that the optimal values of %the primal and dual are equal since the primal problem \eqref{dwsl1} satisfies Slater's condition. % %According to \eqref{z_mu}, we may estimate a dual point by setting $\hat{\boldsymbol{\mu}} = \hat{\bf z} % = {\bf A}\hat{\bf x} - {\bf y}$, but it is not guaranteed to be feasible, {\it i.e.}, it may happen %$\hat{\boldsymbol{\mu}} \notin \mathcal{C}$. %As in \cite{kim2007interior}, from an arbitrary ${\bf x} \in \mathbb{R}^n$, a dual feasible estimate %bound on the suboptimality of ${\bf x}$ can be obtained by %\begin{equation} \label{mu_df} % \hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}} = c \left( {\bf A}{\bf x} - {\bf y}\right), %\end{equation} %where %\begin{equation} \label{constant_c} %c = \min\left\{1, \frac{w_i} %{\left|\left({\bf A}^T{\bf A}{\bf x}\right)_i - \left({\bf A}^T{\bf y}\right)_i \right|}, i=1,\cdots, n \right\} %\end{equation} %It is easy to confirm that $\hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}}$ is dual feasible, {\it i.e.}, % $ \hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}} \in \mathcal{C}$. %Thus, $\mathcal{D}\left( \hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}}\right)$ is a lower %bound on ${\bf p}^*$. % %The difference between the primal objective value of $\bf x$ and the associated %$\mathcal{D}\left( {\boldsymbol{\mu}}\right)$ is called the {\it duality gap}, %denoted by %\begin{equation} \label{duality_gap} %\boldsymbol{\eta} \left({\bf x}, {{\boldsymbol{\mu}}} \right) = \frac{1}{2} \left\| {\bf y} - {\bf A}{\bf x} \right\|_2^2 + %\left\|{\bf w} \odot {\bf x}\right\|_1 - \mathcal{D}\left( {\boldsymbol{\mu}}\right) %\end{equation} %We always have $\boldsymbol{\eta} \left({\bf x}, {{\boldsymbol{\mu}}} \right) \geq 0$ and by weak duality, %$\left({\bf x}, {{\boldsymbol{\mu}}} \right)$ %is no more than $\boldsymbol{\eta}$-suboptimal. %At an optimal point $\left({\bf x}^*, {{\boldsymbol{\mu}}}^* \right)$, %$\boldsymbol{\eta} \left({\bf x}^*, {{\boldsymbol{\mu}}}^* \right) = 0$, namely, strong duality holds. %At a point $\left({\bf x}, \hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}} \right)$, inserting \eqref{mu_df} %with \eqref{constant_c} into \eqref{duality_gap} yields %\begin{equation} \label{duality_gap_mu_df} %\boldsymbol{\eta} \left({\bf x}, \hat{\boldsymbol{\mu}}_{\mbox{\scriptsize df}} \right) = %\left({\bf A}{\bf x} - {\bf y} \right)^T\left( \tfrac{1+c}{2}{\bf A}{\bf x}- \tfrac{1-c}{2}{\bf y} \right) %+ \left\|{\bf w} \odot {\bf x}\right\|_1. %\end{equation} %For any $\boldsymbol{\nu} \in \mathcal{C}$, we have $c = 1$, then %\begin{equation} \label{duality_gap_nu} %\boldsymbol{\eta} \left({\bf x}, {\boldsymbol{\nu}} \right) = %\left({\bf A}{\bf x} - {\bf y} \right)^T\left( {\bf A}{\bf x}\right) %+ \left\|{\bf w} \odot {\bf x}\right\|_1. %\end{equation} \section{Proximity Operator} A fundamental building block for using the DWSL1 norm as a regularizer, since it opens the door to the use of proximal gradient algorithm for solving \eqref{dwsl1}, is its Moreau proximity operator, defined as \begin{equation}\label{POofDWSL1} \prox_{\Omega_{\grave{\bf w}}} \left( {\bf v}\right) = \arg\min_{{\bf x}\in \mathbb{R}^n} \left( \Omega_{\grave{\bf w}} \left({\bf x}\right) + \frac{1}{2} \left\|{\bf x}-{\bf v}\right\|^2\right). \end{equation} In this section, we now show how to compute \eqref{POofDWSL1}, by exploiting the fact that DWSL1 is a generalization of the OSCAR regularizer, extending previous results for OSCAR presented in \cite{zeng2013solving} and \cite{zhong2012efficient}. The following lemma is a simple generalization of Lemma 1 from \cite{zeng2013solving}: \vspace{0.1cm} \begin{lemma} \label{lem:lemma1} Consider ${\bf v} \in \mathbb{R}^{n }$, ${\bf P}({\bf v})$, and ${\bf \grave{v}} = {\bf P(v)\, v}$ as defined above. Let \begin{equation}\label{getmustar} {\bf u}^* = \arg\min_{{\bf u}\in \mathbb{R}^n} \left( \|{\bf \boldsymbol{\pi}} \odot {\bf u} \|_1 + \frac{1}{2} \left\|{\bf u}-{\bf \grave{v}} \right\|^2 \right) \end{equation} where ${\bf \boldsymbol{\pi}}\in \mathbb{R}_+^n$ is a positive weight vector. If \begin{equation}\label{keycondition} |\grave{v}_i| - {\pi}_i \geq |\grave{v}_{i+1}| - {\pi}_{i+1}, \;\; \mbox{for}\; i=1,\dots,n-1, \end{equation} then ${\bf u}^*$ satisfies $|u^*_i|\geq |u^*_{i+1}|$, for $i=1,2,...,n-1$. Moreover, if \eqref{keycondition} is satisfied with ${\bf \boldsymbol{\pi}} = \grave{\bf w} = {\bf P}({\bf v}){\bf w}$, then $\prox_{\Omega_{\grave{\bf w}}} ( {\bf v})= \bigl({\bf P}({\bf v})\bigr)^T {\bf u}^* $. \end{lemma} \vspace{0.1cm} We next describe the proximity operator of $\Omega_{\grave{\bf w}}$, based on the results from \cite{zhong2012efficient} and the analysis in \cite{zeng2013solving}, and leveraging Lemma \ref{lem:lemma1}. Begin by noting that, since $\Omega_{\grave{\bf w}}({\bf x}) = \Omega_{\grave{\bf w}}(|{\bf x}|)$ and, for any ${\bf x},{\bf v} \in \mathbb{R}^n$ and ${\bf u} \in \{-1,\, 0,\, 1\}^n$, \[ \|\mbox{sign}({\bf v})\odot |{\bf x}| - {\bf v}\|_2^2 \leq \| {\bf u} \odot |{\bf x}| - {\bf v}\|_2^2, \] we have that \begin{equation} \label{signfunction} \mbox{sign}\left(\prox_{\Omega_{\grave{\bf w}}} \left( {\bf v}\right) \right) = \mbox{sign}({\bf v}). \end{equation} Consequently, we have \begin{equation}\label{xtarfromx} \prox_{\Omega_{\grave{\bf w}}}({\bf v}) = \mbox{sign}\left({\bf v}\right) \odot \prox_{\Omega_{\grave{\bf w}}}(|{\bf v}|), \end{equation} showing that there is no loss of generality in assuming ${\bf v} \geq 0$, \ie, the fundamental step in computing $\prox_{\Omega_{\grave{\bf w}}}({\bf v})$ consists in obtaining \begin{equation}\label{a_with_v} \begin{split} {\bf a} &= \arg\min_{{\bf x}\in \mathbb{R}_+^n} \left( \Omega_{\grave{\bf w}}\left({\bf x}\right) + \frac{1}{2} \left\|{\bf x}- |{\bf v}|\right\|^2 \right). \end{split} \end{equation} Furthermore, both terms in the objective function in \eqref{a_with_v} are invariant under a common permutation of the components of the vectors involved; thus, denoting \begin{equation}\label{b_with_vsort} {\bf b} = \arg\min_{{\bf x}\in \mathbb{R}_+^n } \left( \Omega_{\grave{\bf w}}\left({\bf x}\right) + \frac{1}{2} \left\|{\bf x}- |{ \grave{\bf v}}|\right\|^2 \right), \end{equation} where (as defined above) $ \grave{\bf v} = {\bf P}({\bf v})\;{\bf v}$, allows writing \begin{equation} \label{a_and_b} {\bf a} = \bigl({\bf P}({\bf v})\bigr)^T{\bf b}, \end{equation} showing that there is no loss of generality in assuming that the elements of ${\bf v}$ are sorted in decreasing magnitude. As shown in the following theorem, ${\bf b}$ has several important properties. \vspace{0.2cm} \begin{theorem} \label{the:theorem1} Letting ${\bf b}$ be as defined in \eqref{b_with_vsort}, we have that \begin{itemize} \item [(i)] For $i=1,2,...,n-1$, $b_i\geq b_{i+1}$; moreover, \[(|\grave{v}_p| = |\grave{v}_{q}|) \Rightarrow ( b_p= b_q). \] \item [(ii)] The property of ${\bf b}$ stated in (i) allows writing it as \[{\bf b} = [{ b}_1 ... { b}_n]^T = \left[{ b}_{s_1}... { b}_{t_1}...{ b}_{s_j}... { b}_{t_j}... { b}_{s_l}...{ b}_{t_l} \right]^T, \] where $b_{s_j} = \cdots = b_{t_j}$ is the $j$-th group of consecutive equal elements of ${\bf b}$ (of course, ${s_1} = 1$ and $t_l = n$) and there are $l \in \{1,.., n\}$ such groups. For the $j$-th group, the common optimal value is \begin{equation}\label{average} { b}_{s_j}=\cdots={ b}_{t_j} = \max \left\{{ \bar{v}}_j - \bar{w}_j, 0 \right\} \end{equation} where \begin{equation} \label{zbarandwbar1} { \bar{v}}_j = \frac{1}{\vartheta_j}\sum_{i = s_j}^{t_j}|\grave{v}_{i}|, \end{equation} is the $j$-th group average (with ${\vartheta_j} = t_j - s_j + 1$ denoting its number of components) and \begin{equation} \bar{w}_j = \frac{1}{\vartheta_j} \sum_{i = s_j}^{t_j} \grave{w}_{i}.\label{zbarandwbar2b} \end{equation} \item [(iii)] For the $j$-th group, if ${ \bar{v}}_j - \bar{w}_j \geq 0 $, then there is no integer $r \in \left\{s_j,\cdots,t_j-1\right\}$ such that \[ \sum_{i=s_j}^r |\grave{v}_i|- \sum_{i=s_j}^r \grave{w}_i > \sum_{i=c+1}^{t_j} |\grave{v}_i|- \sum_{i=c+1}^{t_j} \grave{w}_i. \] Such a group is called {\rm coherent}, since it cannot be split into two groups with different values that respect (i). \end{itemize} \end{theorem} \vspace{0.1cm} Theorem \ref{the:theorem1} is a generalization of Theorem 1 in \cite{zeng2013solving}, and Theorem 1 and Propositions 2, 3, and 4 in \cite{zhong2012efficient}, where an algorithm was also proposed to obtain the optimal ${\bf b}$. That algorithm equivalently divides the indices of $|{\bf \grave{v}}|$ into groups and performs averaging within each group (according to \eqref{zbarandwbar1}), obtaining a vector that we denote as ${\bf \bar{v}}$; this operation of grouping and averaging is denoted as \begin{equation}\label{groupandaverage} \bigl(\bar{\bar{{\bf v}}},\bar{\bar{{\bf w}}}\bigr) = \mbox{GroupAndAverage}\left(|{\grave{\bf v}}|,{\bf w}\right), \end{equation} where \begin{equation}\label{newzsort} \bar{\bar{{\bf v}}} = [ \underbrace{{ \bar{v}}_1,\ldots,{ \bar{v}}_1}_{\vartheta_1 \;\mbox{\footnotesize components}} \ldots \underbrace{{ \bar{v}}_j\ldots{ \bar{v}}_j}_{\vartheta_j \;\mbox{\footnotesize components}} \ldots \underbrace{{ \bar{v}}_l\ldots{ \bar{v}}_l}_{\vartheta_l \;\mbox{\footnotesize components}} ]^T, \end{equation} and \begin{equation}\label{newwsort} \bar{\bar{{\bf w}}} = [ \underbrace{{ \bar{w}}_1,\ldots,{ \bar{w}}_1}_{\vartheta_1 \;\mbox{\footnotesize components}} \ldots \underbrace{{ \bar{w}}_j\ldots{ \bar{w}}_j}_{\vartheta_j \;\mbox{\footnotesize components}} \ldots \underbrace{{ \bar{w}}_l\ldots{ \bar{w}}_l}_{\vartheta_l \;\mbox{\footnotesize components}} ]^T, \end{equation} with the $\bar{v}_j$ as given in \eqref{zbarandwbar1} and the $\bar{w}_j$ as given in \eqref{zbarandwbar2b}. Finally, ${\bf b}$ is obtained as \begin{equation}\label{equivalentform}{\bf b} = \max(\bar{\bar{{\bf v}}}-\bar{\bar{\bf w}}, 0 ). \end{equation} The following lemma, which is a simple corollary of Theorem \ref{the:theorem1}, indicates that condition \eqref{keycondition} is satisfied with $\boldsymbol{\pi} = \bar{\bar{\bf w}}$. \vspace{0.1cm} \begin{lemma} \label{lem:lemma2} Vectors $\bar{\bar{{\bf v}}}$ and $\bar{\bar{\bf w}}$ satisfy \begin{equation} \bar{\bar{v}}_i - \bar{\bar{w}}_i \geq \bar{\bar{v}}_{i+1} - \bar{\bar{w}}_{i+1}, \;\;\mbox{for}\; i=1,\dots,n-1. \end{equation} \end{lemma} \vspace{0.1cm} We are now ready to give the following theorem for $\mbox{prox}_{\Omega_{\grave{\bf w}}} ( {\bf v})$, (which generalizes Theorem 2 from \cite{zeng2013solving}): \vspace{0.1cm} \begin{theorem} \label{the:theorem2} Consider ${\bf v} \in \mathbb{R}^{n }$ and a permutation matrix ${\bf P}({\bf v})$ such that the elements of ${\grave{\bf v}} = {\bf P}({\bf v})\, {\bf v}$ satisfy $|\grave{v}_i| \geq |\grave{v}_{i+1}|$, for $i=1,2,...,n-1$. Let ${\bf a}$ be given by \eqref{a_and_b}, where ${\bf b}$ is given by \eqref{equivalentform}; then $\prox_{\Omega_{\grave{\bf w}}} ( {\bf v}) = {\bf a}\odot \sign({\bf v})$. \end{theorem} \vspace{0.2cm} The fact that the proximity operator of DWSL1 can be efficiently obtained makes solving \eqref{dwsl1} possible via state-of-the-art proximal gradient algorithms, such as FISTA \cite{beck2009fast}, TwIST \cite{bioucas2007new}, or SpaRSA \cite{wright2009sparse}, or by using the alternating direction method of multipliers (ADMM) \cite{boyd2011distributed} or the split-Bregamn method (SBM) \cite{goldstein2009split}. The convergences of the aforementioned algorithms are guaranteed by their own convergence results, since the proximity operator stated above is exact. \Section{Conclusions} \label{sec:conclusions} In this short paper, we have presented some results concerning the so-called {\it decreasing weighted sorted $\ell_1$ norm} (DWSL1), which is an extension of the OSCAR ({\it octagonal shrinkage and clustering algorithm for regression} \cite{bondell2007simultaneous}) that are fundamental building blocks for its use and analysis as a regularizer. Namely, after showing that the DWSL1 is in fact a norm, we have derived its dual norm and presented its Moreau proximity operator. % ------------------------------------------------------------------------- \bibliographystyle{IEEEtran} \bibliography{bibfile_spl} \end{document} }

72 — 1404.4314

\caption{Examples of the application of rule 1 (above) and 2 (below), showing the patterns above the sentence and the added dependencies in {\color{red} red} below the text. Dependencies in {\color{blue} blue} will be removed. \label{fig:rules}}

73 — 1404.4651

\caption{Additional \aastex\symbols}

74 — 1404.5318

\caption{Validation of the osmotic pressure law. $\Delta P$ is represented as a function of theoretical $\Pi=icRT$ for intrusion ($o$) and extrusion ($\meddiamond$). van 't Hoff law (- - -) and a simulation from ref \cite{Luo2010} (\textcolor{Fuchsia}{$\relbar$}) are also reported. Top left frame is a magnification of the curve.}

75 — 1404.5344

\caption{despeckling results of SAR-BM3D \cite{ParrilliPAV12} and our approach. {\color{blue}{Our}} results are marked with {\color{blue}{blue}} color (with{\color{blue}{ \eqref{newmodel}}}). The results are reported with PSNR and SSIM values.}

76 — 1404.5680

\caption{\textbf{a}, Vibrational branching in SrF. Solid upward lines denote transitions driven by the MOT lasers. Spontaneous decays from the A$^2\Pi_{1/2}(v'=0)$ state (solid wavy) and A$^2\Pi_{1/2}(v'=1,2)$ states (dashed wavy) are governed by the vibrational branching fractions $b_{0v}$, $b_{1v}$, and $b_{2v}$, as shown. \textbf{b} Optical addressing scheme for the SrF MOT presented and discussed in the main text. \textbf{c}, Energy levels of the $X^2\Sigma_{1/2}(v=0,N=1)$ state versus $B$. Energy levels are labeled by their $m_F$ value with $m_F=2$ {(\color{red}\LARGE\textbf{-}\normalsize\color{black})}, $m_F=1$ {(\color{orange}\LARGE\textbf{-}\normalsize\color{black})}, $m_F=0$ {(\color{green}\LARGE\textbf{-}\normalsize\color{black})}, $m_F=-1$ {(\color{blue}\LARGE\textbf{-}\normalsize\color{black})} , $m_F=-2$ {(\color{Fuchsia}\LARGE\textbf{-}\normalsize\color{black})}.}

\caption{\textbf{a}, Optical addressing scheme for the SrF MOT. Relevant energy levels are shown for a positive $B$-field. The $\mathcal{L}_{00}$ laser (\textcolor{Red}{\protect\rule[2pt]{.1cm}{2pt}}\;\textcolor{Red}{\protect\rule[2pt]{.1cm}{2pt}}\;\textcolor{Red}{\protect\rule[2pt]{.1cm}{2pt}}) primarily addresses the $|J\!\!=\!\!3/2,F\!\!=\!\!2\rangle$, $|J\!\!=\!\!3/2,F\!\!=\!\!1\rangle$, and $|J\!\!=\!\!1/2,F\!\!=\!\!0\rangle$ states, while the $\mathcal{L}_{00}^{\dagger}$ laser (\textcolor{Orange}{\protect\rule[2pt]{.1cm}{2pt}}\;\textcolor{Orange}{\protect\rule[2pt]{.1cm}{2pt}}\;\textcolor{Orange}{\protect\rule[2pt]{.1cm}{2pt}}) addresses the $|J\!\!=\!\!1/2,F\!\!=\!\!1\rangle$ state. Arrows show $\sigma^{-}$ transitions (\textcolor{Red}{\protect\rule[0pt]{.4cm}{4pt}}) driven by the $\mathcal{L}_{00}$ laser light and $\sigma^{+}$ transitions (\textcolor{Orange}{\protect\rule[0pt]{.4cm}{4pt}}) driven by the $\mathcal{L}^{\dagger}_{00}$ laser light. Transitions from $|J\!\!=\!\!3/2,F\!\!=\!\!1\rangle$, $|J\!\!=\!\!1/2,F\!\!=\!\!0\rangle$, and $|J\!\!=\!\!1/2,F\!\!=\!\!1\rangle$ are marked with transparent arrows for clarity; each line width is proportional to the transition strength. The lasers are drawn at the ground state energy with which they would be resonant; hence, all laser frequencies are red-detuned to the primary $J,\!F$ sublevel they address. This set of detunings is only employed for the primary trapping lasers; all repump lasers are tuned to the field-free resonance. \textbf{b}, SrF MOT experiment schematic showing the MOT (\textcolor{Red}{\protect\rule[1pt]{.4cm}{2pt}}) and slowing (\textcolor{green}{\protect\rule[1pt]{.4cm}{2pt}}) laser beam paths. The line widths indicate beam diameters, and the grey arrows illustrate the default magnetic field gradient. The $\lambda/4$ waveplates and mirrors used to create the vertical MOT beam (dashed line) are not shown.}

\caption{\textbf{a}, MOT cloud response to rapid displacement of the trap center. Shown are MOT LIF images, each averaged over 1600 pulses (top), 2D Gaussian fits (middle), and the extracted radial position (bottom) as a function of time. (In these images, the radial axis is oriented vertically.) The fit is to the motion of a damped harmonic oscillator (see main text). Zero is set at the position of the MOT with no displacement. \textbf{b}, Free expansion of the MOT following release. Shown are MOT LIF images, each averaged over 2000 pulses (top), 2D Gaussian fits (middle), and measured MOT radial (\footnotesize{\color{black}$\blacksquare$}\normalsize) and axial ({\color{red}$\bullet$}) widths vs. free expansion time (bottom); see main text for fit function. In both \textbf{a} and \textbf{b}, images are rescaled to the maximum value at each time. \textbf{c}, LIF in the trapping region vs. time $t$ for MOT with (\tiny{\color{red}$\blacksquare$}\normalsize) and without (\footnotesize{\color{purple}$\bullet$}\normalsize) the $\mathcal{L}_{32}$ repump laser, for untrapped molasses (\small{\color{black}$\blacktriangledown$}\normalsize), and for damping/anti-restoring (\small{\color{blue}$\blacktriangle$}\normalsize) configurations (see Supplementary Information). Overlaid are single exponential fits. Inset: MOT lifetime vs. MOT laser beam diameter. All error bars show the $\pm 1\sigma$ confidence interval.}

77 — 1404.6046

\caption{ \label{fig:sred} % (Color online) Total reduced cross section \sred\versus reduced energy\ered\for many\al -induced reactions \cite{Mohr13}, compared to the recent \siii \rap \clvi\\cite{Bow13} and \siii \ran \arvi\cross sections; the latter is calculated from the ratio$r$ in Fig.~\ref{fig:ratio}. % }

78 — 1404.6232

\caption[]{a) (left) Azimuthal correlation $C(\Delta\phi)$ (\fullsquare\;) of$h^{\pm}$ with $1\leq p_{T_a}\leq 2.5$ GeV/c from a ``high $p_T$'' trigger $h^{\pm}$ with $2.5\leq p_{T_t}\leq 4$ GeV/c in Au+Au central (0-5\%) collisions at $\sqrt{s_{NN}}=200$ GeV~\cite{PXhpairPRL98}. The solid line is the correction for $v_2$ which is subtracted to give the jet correlation function $J(\Delta\phi)$ ({\color{Red}\;\fullcircle}); b)(right) $C(\Delta\phi)$ for like sign $h^{\pm}$-pairs with $0.2<p_{T_1}, p_{T_2} <0.4$ GeV/c for both particles~\cite{JTMQM06}.}

\caption[]{a) (left) PHENIX measurement~\cite{ppg100} of mid-rapidity \Et distributions for p-p (\opensquare ) and d+Au ($\opencircle\!$) at \sqsn=200 GeV with calculations of the d+Au spectrum from fits to the p-p data (lines) based on the AQM (color-strings) and the number of constituent-quark participants (NQP). b) (right) Au+Au \Et distribution and NQP calculation. Systematic uncertainties are shown by dashed lines in both (a) and (b). \label{fig:PXNQP}}

79 — 1404.7301

\caption{Power plots for 4 procedures plotted against the number of points sample per curve ($M$). Method L2 is $\bolds{\circ}$, PC is \protect\includegraphics{692i01.eps}, PC5 is \protect\includegraphics{692i02.eps}, and MV is \protect\includegraphics[raise=-1pt]{692i03.eps}. The left, middle and right panels correspond to the linear, normal c.d.f. and sinusoidal signals.}

80 — 1404.7815

\caption{Dispersion curves for \Web=760 and \Rey=1060. Growth rate vs wavenumber at various times after impact for (a) the theory of \citet{Agbaglah13} and (b) inviscid Rayleigh-Plateau (\dashed{0.5mm}{0.5mm}) and Rayleigh-Taylor (\solid{2.5mm})). }

81 — 1404.7827

\caption{$L(\cdot,\cdot)$ operator is a linear operation which depends on the channel coefficients \textcolor{c3}{not known at Tx}. Therefore, the $L$ operator changes over the time. By using joint encoding over the states together with symbol extension, we can transmit 29 symbols over 14 channel uses.}

82 — 1405.0011

\caption{ \label{fig:filters} The Frontier Fields are being observed with these seven \Hubble\ACS and WFC3/IR filters. Response curves are plotted versus wavelength ($\lambda$) with the corresponding \Lya\redshift ($z$) given along the top axis ($\lambda = 0.1216\mu$m $(1+z)$). F105W and F140W are offset vertically for clarity. Dots mark the effective ``pivot'' wavelengths \citep{TokunagaVacca05} of the filters. % lam = 0.1216 * (1+z) % z = lam / 0.1216 - 1 %The response curves of F105W and F140W are offset vertically for clarity. }

\caption{ \label{fig:volumes} Cumulative area and corresponding co-moving volume at $z \sim 9$ as a function of magnification for each cluster according to the CATS (left) and Zitrin-LTM (center) models. % On both plots we also show both model predictions for the total volumes lensed by all six clusters. % A corresponding plot for the Sharon models is presented in T.~Johnson et al.~(2014, in preparation). % At right, we compare the total volume predictions for all three models to highlight their strong agreement. %The total volumes lensed by all six clusters according to both models %are shown on both plots to highlight the strong overall agreement. % The full survey will yield \~28 arcmin\squared\(\~50,000 Mpc\cubed\at$z \sim 9$) in the 6 blank WFC3/IR fields and \~5 arcmin\squared\(\~9,000 Mpc\cubed\at$z \sim 9$) of source plane search area in the 6 lensed WFC3/IR fields. % delens/volumes.py: % 5.62 CATS % 4.91 Sharon % 4.20 Zitrin-LTM % These are upper limits as we do not account for area lost due to foreground objects. % The plots also show, for example, that in the lensed fields, a total of \~1,000 Mpc\cubed\$z \sim 9$ source plane area should be magnified by a factor of 6 (\~2 magnitudes) or greater.}

\caption{ \label{fig:lensLFs} Optimistic numbers of lensed images of $z \sim 9$ galaxies expected for each Frontier Field. We adopt an optimistic evolution of the \cite{Bradley12b} LF from $z = 8$ to 9 (\S\ref{sec:fieldLF}) and lens this through every lens model submitted, yielding numbers of lensed images per unit magnitude and unit redshift within the WFC3/IR FOV (4.65 arcmin\squared) as a function of observed magnitude. % Incompleteness and contamination are not considered here; these must be estimated and accounted for in any search for high-redshift galaxies. }

\caption{ \label{fig:seds_FFC1} SEDs of two confident $z \sim 8$ candidates lensed by Abell 2744 based on deep WFC3/IR and shallow ACS imaging (see Table \ref{tab:FFC1}). % We plot the observed fluxes as filled green circles (or triangles for non-detections) with 1\sig\vertical error bars.% Horizontal error bars mark the filter wavelength ranges. % We plot our best fitting spectral template in blue as well as the best $z < 4$ template in red. We integrate these spectra through the \HST\filters to plot model fluxes as open squares. Both candidates were also identified by\cite{Atek14} and \cite{Zheng14}, and FFC1-2508-2497 (right) was studied further by \cite{Laporte14}. Deep ACS imaging is upcoming for these candidates. %We expect upcoming deep ACS imaging to further improve the confidence of these candidates. }

83 — 1405.0500

\caption{Illustration of the \predis\construction in the semiring$(\Rset_+, +, \times, 0, 1)$. For each state $(q, s)$ of the result, the subset $s$ is explicitly shown. $q$ is the state of the first pair in $s$ shown. The weights are rational numbers, for example $1/11 = \frac{1}{11} \approx .091$.}

84 — 1405.0516

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Top panel:} shows a 3-d visualization of a galaxy of stellar mass $4.7\times 10^{11}\msun$ at $z=3$. The box has a width of $2.6$ times the virial radius. The (yellow,purple) isodensity surfaces have values $(\sim 10^{-2},\sim 10^{-1})$cm$^{-3}$. {\color{burntorange}\bf Bottom panel:} shows the gas inflow flux in the radial range $(1-1.3)r_v$, in the two-dimensional J-$\cos(\theta)$ phase space. The total gas inflow rate of the galaxy is $388\msun$yr$^{-1}$ (and star formation rate of $254\msun$yr$^{-1}$) and nine significant streams are identified. The top three streams make up $90\%$ of the total inflow rate and are labelled with numbers $(1,2,3)$, with their respective inflow rates being $(211,79,61)\msun$yr$^{-1}$. }

85 — 1405.0698

\caption{({\it a}) Root-mean-square of the spanwise velocity fluctuations during periods without (\textcolor{blue}{---}) and with bursts (\textcolor{red}{---}). ({\it b}) Time series of the plane averaged wall shear stress on the HTS (\textcolor{green}{---}) and WTS (\textcolor{blue}{---}), and volume integrated turbulent kinetic energy (\textcolor{red}{---}).}

\caption{({\it a}) Amplitude $u'$ and ({\it b}) profile at $t=8040$ of the $(\alpha=1, \beta=0)$-mode in the simulation (\textcolor{red}{---}) compared to the growth rate and eigenfunction predicted by the linear stability analysis (\textcolor{blue}{$---$}).}

86 — 1405.0958

\caption{\label{fig:targets} (Color online) Sketch of the targets used in our simulations. In (\textcolor{blue}{a}) is a flat target with 1 $\mu$m pre-plasma. In (\textcolor{blue}{b}) is a target with slab structures on the front, the depth and width of the slabs are 10 $\mu$m and 1 $\mu$m. The spacing between the slabs is 2 $\mu$m. In (\textcolor{blue}{c}) is a target with tower structures on the front, the width of the towers is 1 $\mu$m. As shown in Table \textcolor{blue}{I}, the depth of the towers is usually 10 $\mu$m but is extended to 15 $\mu$m for one simulation. The spacing between towers in both transverse directions is 2 $\mu$m.}

\caption{Details of parameter scan and basic summary of results. The first column indicates the nomenclature for each case. The second column indicates the type of target studied. The third column gives the focal position of the laser relative to the structures - labels A, B, and C correspond to the sketches in the inset to the right. Focus C is used to examine the effect of experimental misalignment. The fourth column indicates the relevant length scale; for the flat this is the pre-plasma scale length and for the slab and towers it is the length of the feature. The peak laser intensity (set by the beam waist, holding all else constant) is given in the fifth column. The sixth column is the conversion efficiency from laser energy to electron energy, counting only those that cross a plane 5 $\mu$m into the target with kinetic energy larger than 1 MeV. The seventh column is the conversion efficiency from laser energy to $\gamma$-ray energy for a gold converter as discussed below in Section \textcolor{blue}{\ref{sec:xray}}.}

\caption{\label{fig:spectra} (Color online) The electron energy distribution for the targets used in our 3D PIC simulations. The naming convention of the legend is consistent with Table \textcolor{blue}{I}. In the high intensity cases (III, V and VI), there is a significant reshaping of the spectrum by the tower structures compared to the flat which produces a much larger number of electrons with energy larger than 50 MeV. The lower intensity case (IV) produces a similar result to the flat target but with higher overall conversion efficiency. }

\caption{\label{fig:ang_distrib_3D} (Color online) Angular distribution of fast electrons ( $>$ 1 MeV). The top graphs are fast electron number distributions (on a color log scale) as a function of kinetic energy and angle $\theta$ with respect to the forward (z) direction. The bottom graphs are fast electron number distributions as a function of momentum direction with respect to the yz and xz planes. The titles of the top graphs indicate the simulation consistent with the naming convention in Table \textcolor{blue}{I}. For the flat target \textcolor{blue}{(I)}, an overall divergence angle of about ${\rm 60}^{\circ}$ is seen. The slab structured target \textcolor{blue}{(II)} is found to have an overall divergence angle of about ${\rm 40}^{\circ}$. The tower structured targets \textcolor{blue}{(III-VI)} show significant narrowing as well as some interesting features. In \textcolor{blue}{III}, the distribution shows two peaks at $\theta_{y}\approx\pm{\rm 4-5}^{\circ}$; for electrons $>$ 100 MeV, each cone angle is about ${\rm 4-5}^{\circ}$. In \textcolor{blue}{IV}, the low intensity case, energy distribution is cooler and the angular peaks disappear due to the reduced strength of the laser. In \textcolor{blue}{V}, the peaks are merged due to the longer length of the tower. In \textcolor{blue}{VI}, the misalignment modifies the angular distribution.}

\caption{\label{fig:Xray_ang_spec} (Color online) The energy \textcolor{blue}{(a)} and angular \textcolor{blue}{(b)} distributions for the radiation generated using Monte Carlo simulations with PIC simulation generated electrons for the conditions described in Table \textcolor{blue}{I}. As expected, the trends are similar to the results for the electron distributions. The high intensity tower structures \textcolor{blue}{III} and \textcolor{blue}{V} show the highest energies. The cutoff energy from \textcolor{blue}{V} is higher than \textcolor{blue}{III} due to longer tower length. Compared to \textcolor{blue}{III}, the misaligned tower \textcolor{blue}{VI} generates lower energy X-rays, but they are still higher than other target types. The low intensity tower structure shows a similar energy distribution as the high intensity flat target, but the angular distribution is narrowed by a factor of 2 at almost all of the energy levels. Other tower targets at high intensities have slightly larger angles than the low intensity tower in its energy range, but much more collimated than the flat. At higher energies, the high intensity towers have further narrower distributions, the FWHM divergence angle is around or smaller than 10$^{\circ}$.}

87 — 1405.0986

\caption{Entanglement detection of $\varrho_\alpha$ \eq{def:rhoalpha} with the help of \eq{eq:Cauchy4}. $E>0$ indicates entanglement. For \textcolor{blue}{$\times$} the measurement basis was optimized for each $\alpha$, whereas for \textcolor{red}{$+$} we used the optimized basis for $\alpha=2$ for all $\alpha$. See the text for further details. \label{fig:rhoalpha}}

88 — 1405.1096

\caption{\label{f-ReducTraj1} (Color online) The \slicePlane\\pSRed, which passes through the {\template} point $\slicep$ and is normal to its group tangent $\sliceTan{}$, intersects all group orbits (dotted lines) in an open neighborhood of $\slicep$. The full \statesp\trajectory$\ssp(\tau)$ (solid black line) and the \reducedsp\trajectory$\sspRed(\zeit)$ (solid green line) belong to the same group orbit $\pS_{\ssp(\zeit)}$ and are equivalent up to a `moving frame' rotation by phase $\theta(\zeit)$. Adapted from \wwwcb. }

\caption{ (Color online) Traveling wave $\REQV{}{1}$ with {\phaseVel} $c = 0.737$ in configuration space: (a) the full \statesp\solution, (b) symmetry-reduced solution with respect to the lab time, and (c) symmetry-reduced solution with respect to the in-\slice\time.\Rpo\$\period{p} = 33.50$ in configuration space: (d) the full \statesp\solution, (e) symmetry-reduced solution with respect to the lab time, and (f) symmetry-reduced solution with respect to the in-\slice\time.}

89 — 1405.1520

\caption{Comparison of two \clasp{} configurations, the \sbs{} solver in all portfolios \iclp{(cf. Subsection~5.3)}, \claspfolio{} 1.0, the \claspfolioTwo{} with \claspredyn{} features, \hydraP{} and \satzillalike{11} approach. %\claspfolioTwo{} consists of \claspredyn{} features, \hydraP{} and \satzillalike{11} approach %and \claspfolioTwo{}(\comp{}) consists of \measp{} features, \isacP{} and \satzillalike{11} approach. The significantly best performances (except VBS) are shown in boldface (according to a permutation test with $100\ 000$ permutations and significance level $\alpha=0.05$). }

90 — 1405.1800

\caption{\it The dilution factor in Eq.~\eqref{dilution_def} as a function of the energy of the overall c.m. frame ($\sqrt{\hat{s}}$) with $m_{H^\pm}=400\rm {~ GeV}$(solid black) and $m_{H^\pm}=600\rm{~GeV}$ (dashed red){\color{red}\cite{Cao:2013ud}}. }

91 — 1405.1922

\caption[Monte-Carlo study of the efficency of the follow-up pipeline.]{ Monte-Carlo study of the efficiency of the follow-up pipeline. The panel (a) shows the percentage of the injected signals without second-order spindown classified as recovered (\textcolor{red}{$-$} $\ \SMC$), and with a non-Gaussian origin ($\textcolor{green}{\times}\ \NGO$), as function of the average $2\mF$ value of the candidate after the original semicoherent Hough-transform search ($\avg{2\mF}_\cand$). The panel (b) shows the percentage of the injected signals with second-order spindown classified as recovered (\textcolor{red}{$-$} $\ \SMC$), and with a non-Gaussian origin ($\textcolor{green}{\times}\ \NGO$), as function of $\avg{2\mF}_\cand$. The error bars account for $95\%$ confidence level. The $90\%$ detection probability is marked with the horizontal dashed line. The vertical dotted line denotes the $2\mF$ threshold used to select candidates in \cite{Aasi:2012fw}.}

92 — 1405.3009

\caption{(Color online) Comparison of the fraction of active nodes $S_a$ and the giant component $G$ after dynamics are terminated in BA networks: theoretical results of all remaining active nodes $S_a$ ({\color{blue}{$\vartriangle$}}), simulation results of $S_a$ ({\color{black}{$+$}}), and giant components of active nodes $G$ ({\color{red}{$\circ$}}). For theoretical results, the turning points (whose abscissas are $\alpha_c$) are marked by solid diamond ($\blacklozenge$).The average degree of BA networks here are $\langle k \rangle = 10$.}

93 — 1405.3164

\caption{\textcolor{CorCol}{The parameters of the GM noise distributions}}

94 — 1405.3253

\caption{\red{ Material source regions for simulated planets. Each pie plot corresponds to a simulated planet and each row to one simulation. Each planet's final orbital position is indicated by its locations along the x-axis and its mass is proportional to the radius of the pie plot cubed. Each colored slice represents material that originated from the region of the disk indicated in the color bar. }}

95 — 1405.3382

\caption{\label{fig:SynScen} \sf Scenarios in MS dataset. ~The sign~\protect\includegraphics[scale=0.5]{abruptsign.jpg} denotes the occurrence of an abrupt drift in the nature of data.}

96 — 1405.4018

\caption{Distribution of the dimensionless spin parameter $\lambda$ in dark matter haloes of different mass at redshift $z \sim 25$ (top panel), $z \sim 20$ (middle panel), and $z \sim 15$ (bottom panel). Crosses indicate the median value of $\lambda$, while the error bars indicate the 20th and 80th percentiles. \MScolor{The dashed vertical lines represent lower mass limit of 1000 particles.} We exclude any bins that contained less than 10 samples. The same holds for Fig.~\ref{virial}.}

\caption{The top panel is the spin distribution for all haloes with \MScolor{$\geq 1000$} particles at $z = 15$. The bottom panel shows the spin distribution for haloes within a mass range of $10^{5\pm 0.2} \,h^{-1} {\rm M}_{\odot}$ (diamonds) and $10^{4\pm 0.2} \,h^{-1} {\rm M}_{\odot}$ (crosses). The symbols represent data from our simulations, and the lines indicate log-normal fits. The vertical solid (dashed) lines represent $\bar{\lambda}$ for our (\citealt{Jan01}) simulation.}

\caption{Mass function of dark matter haloes at $z = 15$ \MScolor{(diamond symbols represent number counts for FOF haloes, when crosses denote those for the main subhalo in a FOF halo)}. The Press-Schechter \citep{PS74} and Sheth-Tormen functions \citep{ST99} are over-plotted.}

\caption{Sphericity $s$ \& triaxiality$T$ for haloes with mass $10^{4\pm 0.2} \, h^{-1} {\rm M}_{\odot}$, $10^{5\pm 0.2}\, h^{-1} {\rm M}_{\odot}$, and $10^{6\pm 0.2} \, h^{-1} {\rm M}_{\odot}$ at $z = 15$. Different lines represent distributions for different groups in each mass range. (Blue dotted lines correspond to halos with $\lambda < \lambda_{1/3}$, dash-dot lines correspond to halos with $\lambda > \lambda_{2/3}$, and black solid lines correspond to all \MScolor{sub}haloes in that mass range.)}

\caption{Relation between formation time and virial ratios. We grouped the haloes in each mass bin into three groups. Early one-third (old haloes), middle one-third (intermediate haloes), and late one-third (young haloes). We plot the median values for early (crosses) and late (diamonds) thirds. The error bars indicate 20th and 80th percentile values for each group. (Error bars are plotted in solid lines for haloes forming earlier, and in dotted lines for haloes forming later. For easy reference, haloes forming later are slightly offset \MScolor{in mass}.) We found that haloes that formed later (young haloes) have higher virial ratios on average due to lack of time for relaxation.}

\caption{\MScolor{Relation between formation time and spin parameter. We grouped the haloes in each mass bin into three groups. We plot the median values with error bars for early (crosses) and late (diamonds) thirds as in Fig.~\ref{t-r}. We found that haloes that formed later (young haloes) have values of spin parameter due to recent mass accumulation.} \MScolor{The dashed vertical lines represent lower mass limit of 1000 particles.}}

\caption{\MScolor{Non-dimensional spin parameter calculated from high-resolution ($2048^3$) simulation and lower-resolution ($512^3$) simulation. Black plus signs denote `main' subhalo with $<= $ 300 particles, green diamonds $>$ 300, $<=$ 1000 particles, blue triangles $>$ 1000, $<=$ 3000 particles, red crosses $>$ 3000 particles in the lower resolution simulation.}}

\caption{\MScolor{Distribution of the dimensionless spin parameter $\lambda$ in dark matter haloes of different mass at redshift $z \sim 25$ (top panel), $z \sim 20$ (middle panel), and $z \sim 15$ (bottom panel) calculated for {\bf FOF haloes}. Crosses indicate the median value of $\lambda$, while the error bars indicate the 20th and 80th percentiles. The dashed vertical lines represent lower mass limit of 1000 particles.}}

97 — 1405.4126

\caption{ Phase description extracted directly from rhythmic signals in a system of two mutually coupled van der Pol oscillators \red{\cite{SM}.} %$\dot{x}_1 = y_1 + K(x_2-x_1) + \xi_{x,1}(t), %\dot{y}_1 = \epsilon_1(1-x_1^2)y_1 - x_1 + Kx_2^2 y_2 + \xi_{y,1}(t), %\dot{x}_2 = y_2 - K x_1^2 y_1 + \xi_{x,2}(t), %\dot{y}_2 = \epsilon_2(1-x_2^2)y_2 - x_2 + K x_1 y_1^2 + \xi_{y,2}(t)$, %with $\langle \xi_{a,i}(s)\xi_{b,j}(t) \rangle = \sigma^2\delta_{ij}\delta_{ab}\delta(s-t),\epsilon_1 = 0.3, \epsilon_2=0.7, K=0.01$ and $\sigma = 0.03$. %From each oscillator, \red{From the two state variables of each oscillator,} only the variable $y_i(t)$ is observed as a signal, as depicted in (a). (b) The recorded signals $y_i(t)$ (left graphs) are transformed into phase time series $\phi_i(t)$ (right graphs). Both $y_i(t)$ and $\phi_i(t)$ fluctuate slightly, owing to the interactions and noise, though the fluctuations are difficult to discern here. Typical data taken over approximately five oscillation cycles are plotted. (c) Posterior probability density distribution of $\hat{\omega}_2$ calculated for three observation durations\red{.} %: %20 (light gray), 160 (dark gray), and 1000 (black) oscillation cycles. \red{The dashed vertical line represents the theoretical value.} %It is seen that as this duration increases, the peak of the posterior %density distribution becomes sharper (greater precision), and the position of the peak %becomes closer to the theoretical value, represented by the dashed vertical line (greater accuracy). (d) Log marginal likelihood for various values of $M_i$ in the case of the data measured with 1000 oscillation cycles. Following the Bayesian model selection, we choose the value $M_i$ giving the largest marginal likelihood. The log marginal likelihoods for $\hat{\Gamma}_{12}$ (red solid curve) and $\hat{\Gamma}_{21}$ (blue dashed curve) have maximum values at $M_1^* =1$ and $M_2^* =3$, respectively. (e) Posterior distributions for the deterministic terms $\omega_i+\Gamma_{ij}(\Delta \phi)$. The blue curves and the light blue regions represent the mean and 95\% confidence interval of the distribution, respectively. The black curves are those obtained from the phase reduction theory, and the numerous gray dots denote data points. The graphs in each column display the results obtained using the data measured with the three different time durations\red{.} %, as indicated at the tops of the graphs. (f,g) Here, it is seen that even if different types of dynamical variables (i.e., \red{$x_1$ and $y_2$}) are used as the signals (panel f), the phase coupling functions can still be reliably retrieved, except for an inevitable uncertainty in the phase shift (see panel g and the main text). For comparison, the dashed curves represent the theoretical curves with the correct phase relationship. }

98 — 1405.4279

\caption{\label{fig:overview}PES $\si(\Delta E)$ in the vicinity of the 2s2p resonance as a function \textcolor{\green}{of} XUV excitation time $t_0$ ($l=1$ partial wave). IR peak intensity $2\times10^{12}W/cm^2$. Solid line: IR field, curved fine-dashed lines: $2n\pi\hbar/|t_0|, n=1,2,3$, closely follow interference maxima. Negative $t_0$ correpsond to the XUV preceding IR. Dashed vertical lines indicate the lineout times of Fig.~\ref{fig:lineout}. }

\caption{\label{fig:tas} TAS at the $2s4p$ resonance. Results are shown for (a) XUV only, (b) a \textcolor{\green}{$7\,fs$} FWHM IR pulse with peak intensity $I=2\times10^{12}W/cm^2$ reached at arrival of the XUV pulse, and (c) the peak IR intensity \textcolor{\green}{$5\,fs$} after the XUV. Dots: numerical results; green lines: fit of a Fano profile with complex $q$ (arrow marks the local minimum), solid red lines: Fano profile with real $q$ according \cite{Ott2013} and additional offset indicated by the dashed red line. }

99 — 1405.4293

\caption{\label{fig:JS} The integral (left) and differential (right) jet shapes of {\color{blue}{quark}} and {\color{red}{gluon}} jets of size $R=0.3$ in proton-proton collisions, plotted as an illustration of their differences. Jet shape contains the information about the transverse energy distribution inside a jet. On average, quark jets are more localized whereas gluon jets are more spread out.}

\caption{\label{fig:event}Schematic event topology of $N$-jet production with {\color{blue}{collinear}} and {\color{red}{soft}} radiation. Jets are reconstructed using a jet algorithm with a parameter $R$. The energy $E_r$ inside a cone of size $r$ in $J_1$ is measured, as well as its transverse momentum $p_T$ and rapidity $y$. An energy cutoff $\Lambda$ outside the jets is imposed to ensure the $N$-jet configuration.}

\caption{\label{fig:fixresum} The integral (left) and differential (right) jet shapes for {\color{blue}{quark}} and {\color{red}{gluon}} jets reconstructed using the anti-$\rm k_T$ algorithm with $R=0.3$, with a {\sl fixed} jet energy $E_J=100$ GeV plotted as an illustration. The dashed lines are the SCET calculations at leading-order (LO), whereas the solid lines are the ones at next-to-leading logarithmic order (NLL). }

100 — 1405.4342

\caption{(Color online) Cascading failures of coupled ER random graphs under random failure. The network size $N_A = N_B = 1000$ and average degree $\langle k_A \rangle = \langle k_B \rangle = 6$. The relative size of the giant component $G$ is shown as a function of the tolerance parameter $\alpha$ with different types of coupling patterns, namely, assortative coupling (\textcolor{magenta}{$\blacksquare$}), random coupling (\textcolor{blue}{$\blacktriangle$}) and disassortative coupling ($\ast$). Cascading failures of the single ER random graph under random failure (\textcolor{red}{$\blacklozenge$}) and the single BA scale-free network under intentional attack (\textcolor{green}{$\bullet$}).}

\caption{(Color online) Cascading failures of coupled ER random graphs under intentional attack. The network size $N_A = N_B = 1000$ and average degree $\langle k_A \rangle = \langle k_B \rangle = 6$. The relative size of the giant component $G$ is shown as a function of the tolerance parameter $\alpha$ with different types of coupling patterns, namely, assortative coupling (\textcolor{magenta}{$\blacksquare$}), random coupling (\textcolor{blue}{$\blacktriangle$}) and disassortative coupling ($\ast$). Cascading failures of the single ER random graph (\textcolor{red}{$\blacklozenge$}) and BA scale-free network (\textcolor{green}{$\bullet$}) under intentional attack.}

\caption{(Color online) Cascading failures of coupled BA scale-free networks under random failure. The network size $N_A = N_B = 1000$ and average degree $\langle k_A \rangle = \langle k_B \rangle = 6$. The relative size of the giant component $G$ is shown as a function of the tolerance parameter $\alpha$ with different types of coupling patterns, namely, assortative coupling (\textcolor{magenta}{$\blacksquare$}), random coupling (\textcolor{blue}{$\blacktriangle$}) and disassortative coupling ($\ast$). Cascading failures of the single BA scale-free network under random failure (\textcolor{red}{$\blacklozenge$}) and intentional attack (\textcolor{green}{$\bullet$}). }

\caption{(Color online) Cascading failures of coupled BA scale-free networks under intentional attack. The network size $N_A = N_B = 1000$ and average degree $\langle k_A \rangle = \langle k_B \rangle = 6$. The relative size of the giant component $G$ is shown as a function of the tolerance parameter $\alpha$ with different types of coupling patterns, namely, assortative coupling (\textcolor{magenta}{$\blacksquare$}), random coupling (\textcolor{blue}{$\blacktriangle$}) and disassortative coupling ($\ast$). Cascading failures of the single BA scale-free network under random failure (\textcolor{red}{$\blacklozenge$}) and intentional attack (\textcolor{green}{$\bullet$}). }

101 — 1405.5028

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

102 — 1405.5125

\caption{ % Phase-diagram of subaqueous patterns in the %($Re_{2\hmean}$,$\Ga ({2\hmean}/{\Dia})^2$) %space as observed in the pipe experiment of %\citet{Ouriemi2009b} corresponding to different %regimes: `flat bed in motion' %({\color{mygray}\solidsquare}); %`small dunes' ({\color{mygray}$\circ$}); %`vortex dunes' ({\color{mygray}\solidtriup}). % Different regimes of sediment bed patterns obtained in the pipe flow experiment of \citet{Ouriemi2009b}, shown in the parameter plane ($Re_b$,$\Ga ({2\hmean}/{\Dia})^2$): `flat bed in motion' ({\color{mygray}\solidsquare}); `small dunes' ({\color{mygray}$\circ$}); `vortex dunes' ({\color{mygray}\solidtriup}). % % For the purpose of comparison of the results of % the pipe experiment with our % channel configuration, % the pipe Reynolds number in the experiment is % adjusted according to the relation % $Re_{2\hmean} \approx 1.85\pi Re_{pipe}/4$ as % suggested by \citet{Ouriemi2009b}. % For the pipe flow data the Reynolds number $Re_{pipe}$ % $Re_{pipe}=u_bd_{pipe}/\nu$ based upon the pipe diameter $d_{pipe}$ and the bulk velocity $q_f/d_{pipe}$ is used. % The data points in the turbulent channel flow experiment of \citet{Langlois2007a} are indicated by: $D=100\mu m$ ($\vartriangle$); $D=250\mu m$ ($\vartriangleleft$); $D=500\mu m$ ($\vartriangleright$). % The following symbols refer to the present simulations: {\color{black}\solidcircle},~\caseLa; {\color{red}\solidcircle},~\caseLb; {\color{blue}\solidcircle},~\caseTa. }

\caption{ Time evolution of the box averaged turbulent kinetic energy. The simulation of case \caseTa\was initially run on a coarse grid while fixing the particles in space (dashed-line in the plot). The fully developed flow field was then refined to the current grid resolution. Mobile particles (except those at the bottom) were released first at time$t=0$ thus insuring the flow field was fully-developed (in the finer resolution) when the particles were released ($t>0$). }

\caption{Wall-normal profiles of the mean streamwise fluid velocity (black) and total shear stress (blue). (\textit{a}), \caseLa; (\textit{b}) \caseTa. The total shear stress in case \caseTa is the sum of the viscous shear stress ({\color{blue}\dashed}) and the Reynolds shear stress ({\color{blue}\chndot}). The horizontal dashed lines show the offsets corresponding to the mean amplitude of the patterns ($\pm\hat{h}_{b,av}$). Friction velocity \ufric\is determined by fitting a line to the linear region of$\langle \tau \rangle$ and extending it to $y_0$. }

\caption{ (\textit{a}) Time evolution of the mean wavelength of the sediment bed height normalized with the particle diameter. The dashed lines indicate the wavelengths of the second to sixth streamwise harmonics in the current domain. % (\textit{b}) Time evolution of the r.m.s.\sediment bed height. The dashed line shows the fit obtained by\citet[figure~6$a$]{Langlois2007a} for their case with $Re_{2\hmean}=15130$, $\shields=0.099$, $\hmean/\Dia=35$. Note that this fit was obtained for $t\geq740\ubulk/\hmean$ (their first data point is indicated by the symbol `$\vartriangleright$'). % In both graphs solid lines with the following colors correspond to the present cases: {\color{black}\solidthick}~\caseLa, {\color{red}\solidthick}~\caseLb, {\color{blue}\solidthick}~\caseTa. }

\caption{%\revision{}{% Analysis of the volumetric particle flow rate (per unit spanwise length), $\overline{q}_p(x,t)$. (\textit{a})~The time evolution of the streamwise average value, $\langle\overline{q}_p\rangle_x(t)$ in cases \caseLa\(black color) and \caseLb\(red color) is shown with solid lines. The viscous scale $q_v=Ga^2\,\nu$ is used for the purpose of normalization. % The solid lines correspond to the average flow % rate over the entire streamwise period, whereas Additionally, the two dashed lines in each case indicate the respective minimum and maximum values ($\min_x\overline{q}_p$, $\max_x\overline{q}_p$). % (\textit{b})~The same quantity in case \caseTa, using the inertial scale $q_i=u_gD$ for normalization. % (\textit{c})~The value of the time average $\langle\overline{q}_p\rangle_{xt}$ % of $\langle q_p\rangle$ (solid lines in $a,b$), additionally % averaged in time over the final part of the simulations plotted versus the Shields number $\shields$: {\color{black}\solidcircle},~\caseLa; {\color{red}\solidcircle},~\caseLb; {\color{blue}\solidcircle},~\caseTa. Note the different scalings ($q_v$ in the laminar cases, $q_i$ in the turbulent case). The open circles are for featureless bedload transport in laminar flow %\citep{kidanemariam:14a-url}; \citep{kidanemariam:14a}; the dashed line is the fit $\langle\overline{q}_p\rangle_{xt}/q_v=1.66\,\shields^{3.08}$ from that reference. % The solid line is the \cite{wong:06} version $\langle\overline{q}_p\rangle_{xt}/q_i=4.93\,(\shields-0.047)^{1.6}$ of the \cite{meyer-peter:48} formula for turbulent flow. %} }

103 — 1405.5490

\caption{Screenshot of timeline of a Twitter user when \TweetCred browser extension is installed.}

\caption{Data flow steps of the \TweetCred extension and API.}

\caption{CDF of response time of \TweetCred. For 82\% of the users, response time was less than 6 seconds and for 99\% of the users, the response time was under 10 seconds. }

\caption{Summary statistics for the usage of \TweetCred.}

\caption{Feedback given by users of \TweetCred on specific tweets ($n=1,273$).}

104 — 1405.5876

\caption{Top: luminosity distance vs. absolute $B$-band magnitude for all of the spiral galaxies (385) found using the magnitude-limiting selection criteria ($B_{\rm T} \leq 12.9$ and $\delta < 0^{\circ}$). The upper limit absolute magnitude can be modeled as an exponential and is plotted here as the solid \textcolor{blue}{blue} line. The dashed \textcolor{red}{red} rectangle is constructed to maximize the number of galaxies in the volume-limited sample. The limiting luminosity distance and absolute $B$-band magnitude are set to be $25.4$ Mpc and $-19.12$, respectively. Bottom: histogram showing the number of galaxies contained in the box in the top panel as the box is allowed to slide to new positions based on the limiting luminosity distance. Note there is a double peak in the histogram maximizing the sample each at 140 galaxies. The two possible combinations are $D_{\rm L} = 25.4$ Mpc and $\mathfrak{M}_{\rm B} = -19.12$ or $D_{\rm L} = 27.6$ Mpc and $\mathfrak{M}_{\rm B} = -19.33$. We chose to use the former (leftmost peak) because its volume-limiting sample is complete for galaxies with dimmer intrinsic brightness. In total, the two samples differed by only 20 non-mutual galaxies, a difference of $\approx 14\%$. Complete volume-limited samples were computed for limiting luminosity distances ranging from $0.001$ Mpc to $100.000$ Mpc in increments of $0.001$ Mpc.\label{Vol-limit}}

\caption{Luminosity function for the 140 member, volume-limited sample of galaxies obtained from the larger CGS sample. The function is given here (solid \textcolor{blue}{blue} line) in terms of the probability density function, fit to the results of Equation (\ref{LF_Eqn}). The function abruptly stops on the dim end due to our exclusion of galaxies with $\mathfrak{M}_{\rm B} > -19.12$. Superimposed for comparison is the $r$-band luminosity functions of $z \approx 0.1$ galaxies selected from the Sloan Digital Sky Survey (SDSS) for all galaxy types \citep{Blanton:2003} and late types \citep{Bernardi:2013}; illustrated as \textcolor[rgb]{0,0.5,0}{green} dashed and \textcolor{red}{red} dotted lines, respectively. These have all been shifted by $B - r = 0.67$ mag, the average color of an Sbc spiral \citep{Fukugita:1995}, which is roughly the median Hubble type of both the CGS and our derivative volume-limited sample.\label{Lum_Fcn}}

\caption{Location of galaxies on the southern celestial hemisphere. Galaxies with measurable pitch angles are marked with \textcolor{blue}{blue} stars and galaxies with unmeasurable pitch angles are marked with \textcolor{red}{red} diamonds.\label{RA_Alt}}

\caption{Pitch angle distribution (dashed \textcolor[rgb]{0,0.5,0}{green} line) and a probability density function (PDF; solid \textcolor{blue}{blue} line) fit to the data. The pitch angle distribution is a ``binless" histogram that we modeled by allowing each data point to be a Gaussian, where the pitch angle absolute value is the mean and the error bar is the standard deviation. The pitch angle distribution is then the normalized sum of all the Gaussians. The resulting PDF is defined by $\mu = 21.44^{\circ}$, ${\rm median} = 20.56^{\circ}$, $\sigma = 9.85^{\circ}$, ${\rm skewness} = 0.58$, ${\rm kurtosis} = 3.47$, and a most probable pitch angle absolute value of $18.52^{\circ}$ with a probability density value of $\phi = 0.042$ deg$^{-1}$.\label{Pitch_Distribution}}

\caption{Black hole mass distribution (dashed \textcolor[rgb]{0,0.5,0}{green} line) and a PDF (solid \textcolor{blue}{blue} line) fit to the data. The black hole mass distribution is a ``binless" histogram that we modeled by allowing each data point to be a Gaussian, where the black hole mass (converted from pitch angle measurements via Equation (\ref{M-P_Relation})) is the mean and the error bar is the standard deviation. The black hole mass distribution is then the normalized sum of all the Gaussians. The resulting PDF is defined by $\mu = 6.88$ dex $M_{\odot}$, ${\rm median} = 6.94$ dex $M_{\odot}$, $\sigma = 0.67$ dex $M_{\odot}$, ${\rm skewness} = -0.59$, ${\rm kurtosis} = 3.61$, and a most probable SMBH mass of $\log(M/M_{\odot}) = 7.07$ with a probability density value of $\phi = 0.63$ dex$^{-1}$.\label{Mass_Distribution}}

\caption{Top: MCMC sampling of the late-type BHMF (rough {\bf black} line) with the best fit model PDF (solid \textcolor{red}{red} line) surrounded by a $\pm 1\sigma$ error region (\textcolor{gray}{gray} shading). When integrated, the area under the curve yields the number density for the entire volume-limited sample, 4.15 $\times$ $10^{-3}$ $h_{67.77}^3$ Mpc$^{-3}$. The plotted data for the top panel is listed for convenience in Table \ref{Values}. Bottom: MCMC sampling of the late-type BHMF with the best fit model CDF (solid \textcolor{red}{red} line) surrounded by a $\pm 1\sigma$ error region (\textcolor{gray}{gray} shading). The CDF visually depicts the integration of the above PDF in the top panel from $M = 0$ until any desired reference point. Here, $\Phi$ is used to indicate an integrated probability, elsewhere $\phi$ is used to indicate a probability density. The upper asymptote approaches the number density for the entire volume-limited sample, 4.15 $\times$ $10^{-3}$ $h_{67.77}^3$ Mpc$^{-3}$.\label{Monte_Carlo}}

\caption{Contribution by the SMBH mass to the cosmic SMBH mass density (solid \textcolor{red}{red} line) surrounded by a $\pm 1\sigma$ error region (\textcolor{gray}{gray} shading). This plot is proportional to Figure \ref{Monte_Carlo} (top), in that this is the product of the BHMF and the SMBH mass ($\phi M$). When integrated, the area under the curve for this plot yields the SMBH mass density, $\rho = 5.54_{-2.73}^{+6.55}$ $\times$ $10^4$ $h_{67.77}^3$ $M_{\odot}$ Mpc$^{-3}$.\label{Mass_Density}}

\caption{Comparison of the BHMFs generated through MCMC sampling (solid \textcolor{red}{red} line with \textcolor{gray}{gray} shading) and through convolution of the probabilities associated with zero (dashed \textcolor{green}{green} line) and 0.38 dex (dotted \textcolor{blue}{blue} line) intrinsic dispersions.\label{Convolved_BHMF}}

\caption{Comparison between our determination of the BHMF for late-type galaxies with our MCMC fit in \textcolor{red}{red} with a \textcolor{gray}{gray} shaded $\pm1\sigma$ error region, zero intrinsic dispersion (dashed \textcolor{green}{green} line), and 0.38 dex intrinsic dispersion (dotted \textcolor{blue}{blue} line); with those of \citet{Marconi:2004}, depicted by {\bf black} triangles (a late-type BHMF is not provided in \citet{Marconi:2004}, we have merely subtracted their early-type function from their all-type function); \citet{Graham:2007}, depicted by \textcolor{blue}{blue} stars; and \citet{Vika:2009}, depicted by \textcolor{green}{green} hexagons. The BHMF of \citet{Vika:2009} is derived using a relationship between SMBH mass and the luminosity of the host galaxy spheroid, applied to a dust-corrected sample of 312 late-type galaxies from the Millennium Galaxy Catalogue in the redshift range $0.013 \leq z \leq 0.18$. The peak of our BHMF is located at $\log(M/M_{\odot}) = 7.06$, whereas theirs is located at $\log(M/M_{\odot}) = 7.50$. However, \citet{Vika:2009} consider BHMF data for $\log(M/M_{\odot}) < 7.67$ to be unreliable because it is derived from galaxies with $\mathfrak{M}_{\rm B} > -18$, according to their relationship. Note that our entire sample consists of galaxies with $\mathfrak{M}_{\rm B} \leq -19.12$.\label{Late-Types}}

\caption{Visualization of all-type BHMF mass functions generated by the addition of the MCMC PDF of our late-type BHMF with the early-type BHMFs of \citet{Marconi:2004}, \citet{Graham:2007}, and \citet{Vika:2009} represented by {\bf black} triangles, \textcolor{blue}{blue} stars, and \textcolor{green}{green} hexagons, respectively.\label{Plot_2}}

105 — 1405.6765

\caption{The circuit diagram of the process of twice-Hadmard-CNOT attack. Here, $|\psi\rangle_{CAB}$ is a GHZ state in the state $(|000\rangle+|111\rangle)/\sqrt{2}$ shared by Calvin, Alice and Bob. $H$ is Hadmard operation, and \protect\includegraphics{Fig2.eps} represents the controlled-NOT gate in which the top line denotes the control qubit, the bottom line the target qubit.}

106 — 1406.0118

\caption{\small Estimates $\eppsh$ (mean and standard deviation over 10 runs) on the {\tt dome} and {\tt hourglass} data, vs sample sizes $N$ for various noise levels $\sigma$; \textcolor{black}{$d'=2$} is in black and \textcolor{blue}{$d'=1$} in \textcolor{blue}{blue}. In the background, we also show as gray rectangles, for each $N,\sigma$ the intervals in the $\epps$ range where the eigengaps of local SVD indicate the true dimension, and, as unfillled rectangles, the estimates proposed by \cite{Guangliang_ChenLittleMaggioniRosasco:multiscale11} for these intervals. \label{fig:synth-epps}}

107 — 1406.0248

\caption{Reynolds shear stress $-\overline{u'^+v'^+}$ and total shear $\tau_{\rm total}=-\overline{u'^+v'^+}+{\rm d}\overline{u}^+/{\rm d}y^+$ distribution. Blue lines, DNS by Iida \& Nagano\cite{Iida98} at $Re_\tau=80$ (\full) and $Re_\tau=60$ (\broken); symbol (\opensquare), experiment by Niederschulte \etal \cite{Niederschulte90} at $Re_\tau=179$. }

\caption{Contours of two-dimensional two-point correlation coefficient $R_{uu}$ in the ($x,y$)-plane; the reference point is at the mid-height $y_{\textrm{\tiny{ref}}}=0.5\delta$. The direction of the mean flow is from left to right. (a) $Re_\tau=180$; (b) $Re_\tau=80$ (XL). \full, positive correlation; \dashed, negative correlation; the line of $R_{uu}=0$ is not shown here. Contour interval $\Delta R_{uu}$:0.05. }

108 — 1406.0312

\caption{We show the effect of pooling a single descriptor encoding (\textcolor{red}{$\Arrow[->]$} or \textcolor{blue}{$\Arrow[->]$}) with a set of tightly-clustered descriptor encodings (\textcolor{ForestGreen}{$\ThickArrow[->]$}). Two pooled representations are shown: \textcolor{red}{$\Arrow[->]$} + \textcolor{ForestGreen}{$\ThickArrow[->]$} = \textcolor{red}{$\DashedArrow[->,densely dashed]$} and \textcolor{blue}{$\Arrow[->]$} + \textcolor{ForestGreen}{$\ThickArrow[->]$} = \textcolor{blue}{$\DashedArrow[->,densely dashed]$}. With sum-pooling (a), the cluster of descriptors \textcolor{ForestGreen}{$\ThickArrow[->]$} dominates the pooled representations \textcolor{red}{$\DashedArrow[->,densely dashed]$} and \textcolor{blue}{$\DashedArrow[->,densely dashed]$}, and as a result they are very similar to each other. With our GMP approach (b), both descriptors contribute meaningfully, resulting in highly distinguishable pooled representations.}

109 — 1406.0503

\caption{\label{fig:magnetic_breakdown}{\bf Magnetic breakdown.} Closeup of the electron spectral function covering the electron pocket and a hole pocket. In evidence a resonant orbit that involves tunneling through a small gap (magnetic breakdown). It is associated with the peaks in Fig.~{\color{blue}4} of the main article marked with golden bands. $p = 10\%$, $P^x_0 = 0.15$, $P^y_0/P^x_0 = 0.2$, $\delta = 0.3$.}

110 — 1406.1174

\caption{2D projections of dark matter density and stellar light (Johnson-K band, see \textcolor{blue}{Torrey \etal 2014} for details) for two galaxies, one elliptical (upper row) and one disk (lower row), with the relevant radial scales and the corresponding 3D spherically-averaged profiles. Even though the density images depict also subhaloes and satellites, these are not accounted for by the fits and profiles on the right-hand side. The thin gray curves in the rightmost panels represent the stellar profiles of a sample of hand-picked Illustris galaxies which have been classified as elliptical (upper panel) and disk galaxies (lower panel) by visual inspection \citep[corresponding to the red and blue dots of Figures \ref{FIG_STELLARSLOPES} and \ref{FIG_GALAXYPROPERTIES}; see][for details]{Vogelsberger:2014b}.}

111 — 1406.1375

\caption{A piece of the cone $V$ in \eqref{eq:VPyth} with markings at the %full %Pythagorean %orbit $\cO$ %set % of all primitive Pythagorean triples $\bx$% %in \eqref{eq:cOIsP} . Points $\bx%\in\cO $ are marked according to whether the ``area'' %sum of the % hypotenuse % and the even side $f(\bx)=\frac1{12}xy$ is % the square of %a in $\cP_{R}$ with $R\le 3$ (\protect\includegraphics[width=.1in]{DotOrange.pdf}), $R=4$ (\protect\includegraphics[width=.1in]{DotBlue.pdf}), or %a $R\ge5$ (%$\bullet$) \protect\includegraphics[width=.1in]{DotBlack.pdf})% number .}

\caption{A piece of the thin Pythagorean orbit $\cO$ %of Pythagorean triples in \eqref{eq:cOthin}. Points $\bx\in\cO$ are again marked according to whether the ``area'' $f(\bx)=\frac1{12}xy$ is in $\cP_{R}$ with $R\le 3$ (\protect\includegraphics[width=.1in]{DotOrange.pdf}), $R=4$ (\protect\includegraphics[width=.1in]{DotBlue.pdf}), or $R\ge5$ (\protect\includegraphics[width=.1in]{DotBlack.pdf}). }

112 — 1406.1419

\caption{The ``classical'' tresino electromagnetic configuration. Note that tresinos are quantum objects so the distance, on average, between the electrons (in {\color{yellow}{yellow}}) is about $15$\,Comptons ($\lambda_{c}$'s.)}

\caption{A illustration showing a formation collision of a proton (in {\color{red}{red}}) and an $\mathrm{{O}}^{2-}$ ion (in {\color{purple}{purple}}), the latter having two unbound electrons (in {\color{yellow}{yellow}}). A PTM is illustrated in the box (see text for explanation).}

\caption{An illustration of the two possible \em{d-$d^{*}$}\rm\,neutron transfer reactions. Protons (in{\color{red}{red}}), neutrons colorless, and electrons (in {\color{yellow}{yellow}}).}

\caption{An illustration of the deuteron tresino $d^{*}$-$^{3}$He neutron transfer reaction. Protons (in {\color{red}{red}}), neutrons colorless, and electrons (in {\color{yellow}{yellow}}).}

113 — 1406.1467

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Left panel:} shows an example of the formation of a massive galaxy at $z=1$ that is overall determined by the large waves indicated by the blue dashed curve. On top of the large linear wave, there is small-scale linear wave of length $1/8$ of the large one with a fluctuation amplitude 1.5 times larger than the large wave. Both long and short waves are chosen to be sinusoidal for the illustration. The sum of the long and short waves is shown as the black solid curve. The horizontal red dot-dashed line of amplitude value $1.68$ indicates the amplitude of the fluctuation that has collapsed by $z=1$. {\color{burntorange}\bf Right panel:} same as the left panel except the small-scale wave has a fluctuation amplitude 10 times smaller than the large wave. Note that the examples are shown in 1-d but meant to be in 3-d. }

114 — 1406.1967

\caption{\textcolor{blue}{Comparison of scrambled digital nets with small WAFOM and those with large WAFOM for $s = 5$. The top figure shows WAFOM values (in $\log_{10}$ scale) for $m = 1, \ldots, 23$. The other figures show the median of relative errors for the Genz functions for $m = 1, \ldots, 23$. }}

115 — 1406.2035

\caption{ Heatmaps of word representations for 10 animals (top) and 10 countries (bottom) for $M=52$ from \thisapproach{}. \textcolor{BrickRed}{Red} indicates negative values, \textcolor{RoyalBlue}{blue} indicates positive values (darker colors correspond to more extreme values); white denotes exact zero. The $x$-axis shows the original dimension index, we show the dimensions from the most negative (left) to the most positive (right), within each block, for readability. \label{fig:patterns} }

\caption{ Two dimensional projections of the \thisapproach{} (left) and SG (right) word representations using the t-SNE tool \citep{tsne}. Words associated with ``good'' are colored in \textcolor{Blue}{blue}, words associated with ``bad'' are colored in \textcolor{BrickRed}{red}. We can see that in both cases most ``good'' and ``bad'' words are clustered together (in fact, they are linearly separated in the 2D space), except for \emph{poor} in the SG case. See supplementary materials for more examples. \label{fig:plot_as} }

\caption{ Two dimensional projections of the \thisapproach{} (top), SG (middle), and NCE (bottom) word representations using the t-SNE tool \citep{tsne}. Words associated with ``good'' (left) and ``man'' (right) are colored in \textcolor{Blue}{blue}, words associated with ``bad'' (left) and ``woman'' (right) are colored in \textcolor{BrickRed}{red}. The two plots on the top left are the same plots shown in the paper. \label{fig:plot_as_mf} }

116 — 1406.2235

\caption{new\_item\_prediction($\mathbf{a}_{newItem}$)}

117 — 1406.2393

\caption{(Color online) AFM peak intensities as a function of applied field for {\Bc} (a--b) and {\Ba} (c--d) measured at several temperatures. Data taken at base temperature are shown in the upper panel (a,c), in which dotted lines and arrows indicate {\bc2} and inflection points, respectively. Panel (a) shows the derivative as well (\textcolor{red}{${\times}$}). Statistical errors of the intensities are smaller than the symbol size}

118 — 1406.5245

\caption{\protect\includegraphics[scale=0.65]{eqstep}\protect\includegraphics[scale=0.65]{eqfull}}

\caption{\protect\includegraphics[scale=0.3]{AnalytGrowth}\protect\includegraphics[scale=0.325]{gamma_vs_beta_5}}

\caption{\protect\includegraphics[scale=0.3]{betas_a}\protect\includegraphics[scale=0.3]{betas_new2_s6}\protect\includegraphics[scale=0.3]{betas_new2_s5}}

\caption{\protect\includegraphics[scale=0.45]{GK_Map_steps08new}\protect\includegraphics[scale=0.45]{omega_new}}

\caption{\protect\includegraphics[scale=0.3]{omegas_a}\protect\includegraphics[scale=0.3]{omega_hr}}

\caption{\protect\includegraphics[scale=0.45]{GK_Map_steps04new}\protect\includegraphics[scale=0.45]{Gi_effect_four}}

\caption{\protect\includegraphics[scale=0.45]{analytic_Ki_betaless}\protect\includegraphics[scale=0.45]{analytic_Ki_betamore}}

\caption{\protect\includegraphics[scale=0.45]{KiEffect2}\protect\includegraphics[scale=0.45]{Kieffect2_num}}

\caption{\protect\includegraphics[scale=0.45]{KiEffect}\protect\includegraphics[scale=0.45]{Kieffect_num}}

\caption{\protect\includegraphics[scale=0.25]{sketch}}

119 — 1406.5382

\caption{ (Color online) Our iDMRG results for $J_1/J_2=-2.5$ obtained with 300 renormalized basis states ($m=300$). (a) Phase diagrams of the Hamiltonian given by Eq.~\eqref{eq:H} for $\delta = 0$ and $\delta = 0.02$. (b) Order parameters for $\delta = 0.02$. Black, green, and red solid lines are extrapolations of the DMRG to the thermodynamic limit~\cite{UO14}. (c) Variations in $Z_2$ indices [ $\beta(\Theta)$ and $\gamma(\Theta,R^{~}_{2z})$:\textcolor{darkgreen}{+}; $\beta(I)$:\textcolor{red}{$\times$}; $\omega(R^{~}_{2x} R^{~}_{2z})$:\textcolor{blue}{$\bigcirc$} ]. (Definitions are given in the text.) The upper/lower panel shows results obtained by dividing the system at a strong/weak $J_1$ bond. }

120 — 1406.5821

\caption{Comparison of the error of the threshold method (\textcolor{blue}{$+$}), and our improved time resolved method ($\textcolor{red}{*}$) for different total measurement times $t_b$ and constant sub-bin time $t_s=0.1$ms. \label{fig:tresh_time}}

\caption{Long time measurement of the time dependent fluorescence of an average bright ion (\textcolor{red}{+}) and an average dark ion (\textcolor{blue}{$\ast$}) and their simultaneous fit ( lines) for a laser power of $36\mu$W and a beam diameter given by $174\mu$m (measurement laser near 369.5 nm). With a collection efficiency of $\eta_0=3.1 \cdot 10^{-3}$ the average measured photon scattering rates are $R_B=16/$ms for the bright state and $R_D=0.3/$ms for the dark state (see \Sec{sec:generalization} for the definition of the scattering rates). }

\caption{Experimental errors of the threshold (\textcolor{blue}{+})and the time-resolved method (\textcolor{red}{$\ast$}) in comparison to each other and to simulations}

\caption{ Comparison of the experimental data evaluated with the original time-resolved method (\textcolor{red}{*}) that considers only one possible state change from bright to dark, our generalized time-resolved method (\textcolor{blue}{+}) that considers several possible state changes from bright to dark and vice versa, and the threshold method (\textcolor{green}{$\triangle$})}

\caption{Comparison of the relative error (left scale ) of the generalized time-resolved method combined with the $\pi$-pulse method (\textcolor{red}{$*$}) and the threshold method combined with the $\pi$-pulse method (\textcolor{blue}{$+$}), for different total measurement times $t_b$ and fixed sub-bin time $t_s=0.1/3$ms. The detection efficiency $N_R$ for the generalized time-resolved (red stair diagram) and the threshold method (blue bar diagram) correspond to the right scale\label{fig:pi_time_tresh}}

\caption{Relative error (left scale) of the the threshold method combined with the $\pi$-pulse method for different total measurement times $t_b$ and different thresholds $n_c=1$ (\textcolor{blue}{$+$}), $n_c=3$ (\textcolor{red}{$*$}), $n_c=5$ (\textcolor{green}{$\times$}). The red bar diagram displays the detection efficiency $N_R$ for $n_c=3$ (right scale). \label{fig:pi_tresh}}

\caption{Error of the double threshold method for lower limit $n_D=0$ and different upper limits $n_B=1$ (\textcolor{blue}{$\times$}), $n_B=4$ (\textcolor{red}{$+$}), $n_B=10$ (\textcolor{green}{$*$}) for different measurement times $t_b$. The red bar diagram displays the detection efficiency $N_R$ for $n_c=4$ (right scale). \label{fig:tresh2l}}

121 — 1406.6478

\caption{Dynamical Facilitation in the absence of pinning. \textbf{a.} Distribution of instanton times $P_{a}(\Delta t)$ for for $\phi =$ 0.73 ({\color{Red} $\boldsymbol \circ$}), $\phi =$ 0.74 ({\color{Blue} $\boldsymbol \blacktriangle$}), $\phi =$ 0.75 ({\color{Black} $\boldsymbol \blacktriangledown$}), $\phi =$ 0.77 ({\color{Brown} $\boldsymbol \square$}) and $\phi =$ 0.79 ({\color{Green} $\boldsymbol \diamond$}), showing that excitations are localized in time. Inset to \textbf{a.} Representative sub-trajectory of a particle coarse-grained over 2s, shown in black, and the corresponding functional $h_{i}(t,t_{a};a)$ shown in red. The instanton time duration $\Delta t$ and the commitment time $t_{a}$ are marked by dotted lines. The displacement is measured in units of $a = 0.5\sigma_{S}$ and the trajectory has been shifted arbitrarily along the Y-axis to make the rise in displacement coincide with region of non-zero $h_{i}(t,t_{a};a)$. \textbf{b.} The function $F(r;a)$ normalized by its value at $r = \sigma_{L}$ for $\phi =$ 0.74 ({\color{Blue} $\boldsymbol \blacktriangle$}), $\phi =$ 0.75 ({\color{Black} $\boldsymbol \blacktriangledown$}), $\phi =$ 0.77 ({\color{Brown} $\boldsymbol \square$}) and $\phi =$ 0.79 ({\color{Green} $\boldsymbol \diamond$}), showing that excitations are localized in space. \textbf{c-e.} Displacement field of particles for $\phi = $0.73 (\textbf{c}), $\phi =$ 0.75 (\textbf{d}) and $\phi =$ 0.79 (\textbf{e}). Displacements are normalized by the excitation size $a$. Particles coloured in maroon indicate displacements $\geq a$. \textbf{f.} Concentration of excitations $c_{a}$ vs $\phi$. \textbf{g.} Facilitation volume $v_{F}(t)$ for various $\phi$s. The symbols and colors in (\textbf{g}) are same as those in (\textbf{a-b}).}

\caption{Effect of Random Pinning. \textbf{a.} Schematic of the trapping potentials (shown in red) created by the holographic optical tweezers. The underlying image represents a small portion of the field of view for $f_{p}$ = 0.06. The image has been generated by averaging over $\sim 15\tau_{\alpha}$. Pinned particles appear bright due to their low mobility and high overlap with initial positions. \textbf{b.} Structural relaxation time $\tau_\alpha$ for free particles (black) and concentration of excitations $c_{a}$ (red) as a function of the fraction of pinned particles $f_{p}$. \textbf{c.} Facilitation volume $v_{F}(t)$ for $f_{p} =$ 0 ({\color{Black} $\boldsymbol \square$}), $f_{p} =$ 0.03 ({\color{Red} $\boldsymbol \circ$}), $f_{p} =$ 0.06 ({\color{Blue} $\boldsymbol \triangle$}) and $f_{p} =$ 0.12 ({\color{Green} $\boldsymbol \bigtriangledown$}). \textbf{d.} Distribution of string lengths $P(n)$ for $\Delta\tau$ corresponding to the maximum string length and \textbf{e.} String length $n_{s}(\Delta\tau)$ for various $f_{p}$. The colors and symbols in (\textbf{d}) and (\textbf{e}) are identical to those in (\textbf{c}).}

122 — 1406.7042

\caption{2-D Cavity of Sec.~\ref{sec:cavity}: frequency response of the cavity obtained from the results of Yee's FDTD ({\color{black} \st{$\;\;\;\;$}}), the implicit method of~\cite{Subcell} ({\color{red} $--$}) and ADI-FDTD ({\color{magenta} -$-$-}). The proposed method, spatial filtering~\cite{SpatialFiltering}, and the method of Gaffar and Jiao~\cite{DJiao} gave the same results and are depicted with a single curve ({\color{blue} -\,\,-\,\,-}). Yee's FDTD was run below the CFL limit ($s=0.99$). All other methods were run above the CFL limit ($s=4.95$).}

\caption{2-D Cavity of Sec.~\ref{sec:cavity}: relative error on the first 6 resonant frequencies obtained with Yee's FDTD run at $s=0.99$(\st{$\;\bigtriangleup\;$}), the implicit method of~\cite{Subcell} ({\color{red} \st{$\;\Box\;$}}), and ADI-FDTD ({\color{blue} \st{$\;\ominus\;$}}). The proposed method, spatial filtering~\cite{SpatialFiltering}, and the method of Gaffar and Jiao~\cite{DJiao} gave the same results and are depicted with a single curve ({\color{magenta} \st{$\;\Diamond\;$}}). Yee's FDTD was run below the CFL limit ($s=0.99$). All other methods were run above the CFL limit ($s=4.95$). Gold standard: analytical formulas.}

\caption{3-D Cavity of Sec.~\ref{sec:cavity}: relative error on the first 6 resonant frequencies. Yee's FDTD (\st{$\;\bigtriangleup\;$}) at CFL = 0.99 and proposed method at $s=0.99$ ({\color{red} \st{$\;\Box\;$}}), $s=1.98$ ({\color{blue} \st{$\;\ominus\;$}}), and $s=2.97$ ({\color{magenta} \st{$\;\Diamond\;$}}). Gold standard: analytical formulas.}

\caption{Focusing metascreen of Sec.~\ref{sec:propagation}: average amplitude of the electric field E$_y$ on the image plane. Comparison of Yee's FDTD (\st{$\;\bigtriangleup\;$}) at $s=0.99$ against the proposed method at $s=0.99$ ({\color{red} \st{$\;\Box\;$}}), $s=2.97$ ({\color{blue} \st{$\;\ominus\;$}}), and $s=4.95$ ({\color{magenta} \st{$\;\Diamond\;$}}).}

\caption{Microstrip filter of Sec.~\ref{sec:filter}: time domain reflected (top) and transmitted (bottom) waveforms computed with Yee's FDTD (\st{$\;\bigtriangleup\;$}) at $s= 0.99$ and with the proposed method at $s=0.99$ ({\color{red} \st{$\;\Box\;$}}), $s=2.97$ ({\color{blue} \st{$\;\ominus\;$}}), and $s=4.95$ ({\color{magenta} \st{$\;\Diamond\;$}}).}

\caption{Microstrip filter of Sec.~\ref{sec:filter}: S$_{11}$ (top) and S$_{21}$ (bottom) parameters obtained with Yee's FDTD (\st{$\;\bigtriangleup\;$}) at $s= 0.99$ and with the proposed method at $s=0.99$ ({\color{red} \st{$\;\Box\;$}}), $s=2.97$ ({\color{blue} \st{$\;\ominus\;$}}), and $s=4.95$ ({\color{magenta} \st{$\;\Diamond\;$}}).}

123 — 1406.7224

\caption{\coloronline\Event-Chain algorithm for continuous pair interactions. a-d) Evolution of the algorithm for four particles through three subsequent collision events. e) Probability for soft disks that a collision event takes place at a distance$ r $ larger than the cutoff $\rmi{r}{c}$ The vertical dashed line is the cutoff chosen in this work. f) Equation of state with different cutoffs ($N=65\kilo$, $n=6$). \label{figECMC}}

\caption{\coloronline\Phase behavior of$r^{-n}$ soft disks for $n \geq 6$. a) Phase diagram as a function of density $\phi$ relative to the density \phihex\of the pure hexatic at coexistence. The non-monotonic liquid-hexatic coexistence interval vanishes around$n=6$. Symbols match the following figures: bullets are liquid states, filled triangles hexatics, of which downward filled triangles the hexatic at \phihex; empty triangles are solids. Center: Local orientational order parameter $\psi_6$ in $N=259\kilo$ particles, the color code is illustrated to the right. Upper row: b) Liquid phase at $(n, \phi) = (8, 1.193)$ (subset of a $1036\kilo$ configuration) c) hexatic. d) solid. Lower row: Coexistence in the $n=64$ system: e) pure liquid close to coexistence. f) At $\phi = 0.889$, the hexatic and the liquid form stripes. g) At $\phi = 0.898$, a small bubble of liquid remains on a hexatic background of uniform orientation. h) Equations of state with $n$ from $6$ through $1024$ (solid lines correspond to $N=65\kilo$ soft disks, symbols correspond to $N=259\kilo$, \philiq\is the liquid density at coexistence, dashed lines are for Yukawa particles).\label{figALLES} }

\caption{\coloronline\a) Pair correlation function$g(x,y)$ along the $x$ axis in the pure hexatic phase, showing exponential decay for large $x/d$ (ensemble average of $N=259\kilo$ configurations after aligning their global orientational order parameters $\Psi_6$, as in Ref.~\cite{BernardKrauth}). b) Two-dimensional pair correlation function $g(x,y)$ for single configurations at the same parameters. The solid line is the $x$ axis for the left-hand plot. c) Square boxes of side length $40 d$ extracted from the $259\kilo$ configurations: Orientational order is preserved as the positional order is lost (color code for $\psi_6$ as in \Fig{figALLES}{b--g}). \label{figPositionalSRO}}

\caption{\coloronline\a) Orientational correlations$g_6(r)$ close to the liquid-hexatic transition, for several exponents $n$. b) Scaling of the orientational order parameter in subblocks of linear size $L_\text B$ (see Ref.~\cite{BagchiAndersen} for details). The KTHNY hexatic is stable below the bold line of slope $-\frac 14$; short-range order corresponds to the steep bold line of slope $-2$. Bullets are liquid states, triangles hexatics (symbols match those in \Fig{figALLES}). \label{figSix}}

\caption{\coloronline\Yukawa interactions exhibiting first-order (top) and continuous liquid-hexatic transitions. a) Pair correlation function along the$x$ axis, as in \Fig{figPositionalSRO}a. b, c) Interaction potential of the soft disks and the Yukawa particles around $r=\sigma$. \label{figYukawa}}

\caption{Thermodynamic data for $r^{-n}$ soft disks: Pressure $\Pcoex$ at the liquid-hexatic transition; density \philiq\and orientational correlation length\corrsix\of the liquid; density\phihex\and positional correlation length\corrpos\of the hexatic at coexistence. The final column is a lower bound for the melting density (solid-hexatic). The densities are accurate to$\approx 0.5\%$. Pressures are computed using the truncated interaction and are thus low by at most $1.1\%$. The statistical sampling error is a decade smaller (error bars in \Fig{figECMC}f). The correlation lengths are determined from the tail of the respective correlation functions and are subject to large errors of $\pm 10\%$ due to the choice of $\phi$. They are consistent with earlier results for hard disks \cite{BernardKrauth}. % \label{tabData} }

124 — 1407.0661

\caption{Low energy Q integrated (a) and $\hbar\omega$ integrated (b) inelastic neutron scattering intensity of $x= 0$ at 1.5 K ({$\bullet$}), 13 K (\textcolor{blue}{$\circ$}), 100 K (\textcolor{green}{$\diamond$}) and 200 K (\textcolor{red}{$\Box$}). Data were obtained by integrating the spectra shown in Fig.~\ref{iris_cont} over the range 0.5 \(\leq |Q| \leq\) 1.5 \AA$^{-1}$ and 0.1 \(\leq \hbar\omega \leq\) 1.0 meV respectively. For (a) the energy binning is 2 ($\Delta \hbar\omega= 0.02$ meV) compared to Figure~\ref{iris_cont}. The dips observed (at all temperatures) at 0.7 \AA$^{-1}$, $\sim$1.4 \AA$^{-1}$ and $\sim$1.8 \AA$^{-1}$ in (b) are due to variations in detector channel sensitivity (see text for more details). The solid lines shows $|F(Q)|^2$ calculated for Pr$^{3+}$ scaled to the data. Insert shows the same data as in (a) multiplied by the Bose factor.}

125 — 1407.0976

\caption{The exponents, $n$ in Eq.\;\ref{eq:powerlaw}, of power-law fits to the data %given by Eq.~\ref{eq:powerlaw} quantify the energy decay rate. The line is a fit to our data, ({\large\boldmath$\times$}, {\large\boldmath$\mathrlap{\times}+$}, {\large\boldmath$+$}), % new line for which the mean decay rate was %$n = -1.18\pm0.02$. % the error is the standard deviation in the values of n. -this line $n = -1.18$ with a standard deviation of $0.02$. % +this line The 95\% confidence interval for each value of $n$ ($\pm0.9$\%) is given %typically $\pm0.9$\% of its mean value, approximately by the size of the symbols. Various additional data were drawn from the literature as follows: {\large\color{Brown}$\filledstar$} \cite{kurian:2009}, % Kurian and Fransson {\large\color{blue}$\filledstar$} \cite{batchelor:1948}, % Batchelor and Townsend {\large$\filledstar$} \cite{wyatt:1955}, % Wyatt thesis {\large\color{red}$\filleddiamond$} \cite{sirivat:1983}, % Sirivat and Warhaft {\large\color{blue}$\filleddiamond$} \cite{yoon:1990}, % Yoon and Warhaft {\large$\filleddiamond$} \cite{warhaft:1978}, % Warhaft and Lumley {\large\color{ForestGreen}$\filleddiamond$} \cite{comte:1966}, % Comte-Bellot and Corrsin {\large\color{Goldenrod}$\filleddiamond$} \cite{mydlarski:1996}, % Mydlarski and Warhaft $\filledmedsquare$ \cite{sreeni:1980}, % Sreeni et al. {\color{blue}$\filledmedsquare$} \cite{white:2002}, % White et al. {\color{Brown}$\filledmedsquare$} \cite{lavoie:2007}, % Lavoie et al. {\color{ForestGreen}$\filledmedsquare$} \cite{antonia:2003}, % Antonia et al. {\color{red}$\filledmedsquare$} \cite{vanDoorn:1999}, % van Doorn et al. {\color{Goldenrod}$\filledmedsquare$} \cite{bewley:2007}, % Bewley et al. {\footnotesize$\newmoon$} \cite{mohamed:1990}, % Mohamed and LaRue {\footnotesize\color{red}$\newmoon$} \cite{uberoi:1966}, % Uberoi and Wallis {\footnotesize\color{ForestGreen}$\newmoon$} \cite{vanatta:1968}, % Van Atta and Chen {\footnotesize\color{blue}$\newmoon$} \cite{uberoi:1963}, % Uberoi {\footnotesize\color{Goldenrod}$\newmoon$} \cite{kistler:1966}, % Kistler and Vrebalovich $\filledmedtriangledown$ \cite{poorte:2002}, % Poorte et al. $\filledmedtriangleleft$ \cite{makita:1991}, % Makita {\color{red}$\filledmedtriangleright$} \cite{krogstad:2011}, % Krogstad and Davidson {\color{blue}$\filledmedtriangleleft$} \cite{valente:2011}, % Valente and Vassilicos {\color{Brown}$\filledmedtriangleup$} \cite{thormann:2014}. % Thormann and Meneveau Closed symbols mark experiments that employed classical grids, as ours did, whereas open symbols mark those that employed modified grids such as fractal grids. % new sentence. Symbols with thin lines mark active-grid experiments. In some of these experiments, %variations in conditions % -this phrase. modifications to the grids % +this phrase. were made deliberately to elicit changes in the decay \cite{lavoie:2007,thormann:2014}. %The present study was the first to revisit the high Reynolds numbers %of KV, %of Kistler and Vrebalovich \cite{kistler:1966}, %and not only reached higher but did so in a better controlled flow. % moved sentence from text body The inset shows on a linear scale of $Re$ the relative decay exponents as described in the text. %Taken together, % -this phrase. Our data reveal that the rate of decay is invariant with respect to the Reynolds number for $Re \gg 10^4$. }

126 — 1407.0984

\caption{The distribution of angular velocity (in nHz) $\mean\Omega$ in the meridional plane from Runs~A--E. $\mean\Omega$ is computed from $\Omega$ first by the longitudinal average and then the time average over the last few \blue{cycles}. The arrows in the leftmost panel show the co-latitudes at which the latitudinal differential rotation is computed in \Eq{equ:pDRt}. }

\caption{From Run A: (a) The large-scale magnetic field over the whole convection zone $\mean{B}= \brac{\brac{B_r}_{\phi}^2 + \brac{B_{\theta}}_{\phi}^2+\brac{B_\phi}_{\phi}^2}_{r\theta}^{1/2}$ normalized by $B_{\rm eq}$. (b) The latitudinal component of meridional circulation $u_\theta(r,\pm32^\circ)$ (smoothed over 5 months) at $r\approx0.95R_\odot$ (black and red) and $r\approx0.73R_\odot$ (blue and green), (c) azimuthally averaged angular velocity $\mean\Omega(0.95R_\odot,\pm32^\circ)$, \blue{ (d) $\mean\Omega(r,0^\circ)$ at $r=0.73R_\odot$ (red dashed) and $r=0.95R_\odot$ (black). } The dashed (solid) line corresponds to the southern (northern) hemisphere. (e) Radial shear $\Delta_\Omega^{(r)}$, and (f) latitudinal shear $\Delta_\Omega^{(\theta)}$, defined in \Eq{equ:pDRt}, as functions of time. }

127 — 1407.1074

\caption{(Color online) The $J^{XXZ}_{1}$--$J^{XXZ}_{2}$ model on the honeycomb lattice, showing (a) the bonds ($J_{1} \equiv$ ----- ; $J_{2} \equiv \textcolor{blue}{- - -}$) and the two sites (\textcolor{red}{$\bullet$}) A and B of the unit cell; (b) the N\'{e}el planar, N(p), state; (c) the N\'{e}el $z$-aligned, N($z$), state; and (d) the N\'{e}el-II planar, N-II(p), state. The arrows represent the directions of the spins located on lattice sites \textcolor{black}{$\bullet$}.}

128 — 1407.1780

\caption{\label{fig:statistics}(Color online) Plot of $D_{i}$ with rescaled interaction time $\Omega t$ for (a) mode $a,$ (b) mode $b$, and (c) coupled mode $ab$ for $\Omega=10^{2},\,\frac{\Delta}{\Omega}=10^{2},\,\alpha=10,$ and $\beta=2$, and\textcolor{green}{{} }(d) mode $a$ for $\Omega=10^{2},\,\frac{\Delta}{\Omega}=10^{2},\,\alpha=10,$ and $\beta=-2$. Negative regions of the plots in (a)-(d) show antibunching.}

129 — 1407.2241

\caption{Consider a line graph with $n$ nodes and a bag that contains all even numbered nodes. For any monotone crusade, the bag obtained after the first step, that is, after curing one node, has a cut that is proportional to $n$, and therefore the impedance of the bag is proportional to $n$. The impedance of the whole graph is $1$ and a monotone crusade which achieves the minimal width cures nodes one by one starting from the rightmost node. Therefore, the impedance of the bag is much higher than the impedance of the graph. \red{[CAN YOU MAKE THE CAPTION TO BE RIGHT-JUSTIFIED RATHER THAN CENTERED TEXT?]} }

130 — 1407.2289

\caption{Large panel on the left shows the $\log g$ - \teff~ for the entire sample. Open circles represent dwarfs selected from the KIC and filled circles dwarfs from the seismic sample. Squares represent subgiants. Diamonds indicate 16~Cyg~A and B evolutionary status. The upper panels on the right compare the $\log g$ values of \blue{\cite{chaplin2014}} and \blue{\cite{huber2013}} to the KIC values. The bottom panels of the right compare the corresponding \teff~values. }

\caption{Surface gravity from asteroseismology (filled squares and circles) determined by \blue{\citealt{chaplin2014}} and from the \blue{\cite{huber2013}} (open circles) as a function of the 8-hour flicker F$_{8}$, \blue{\citep{bastien13}}. Circles represent dwarfs and squares represent the subgiants. The black dashed line corresponds to the relationship derived by \blue{\cite{bastien13}}. The red symbol at lower left represents the Sun $\log g$ and F$_{8}$. Filled triangle represent KIC~2718678, and KIC~12157617.}

\caption{Color---period diagram for the seismic sample for which there are {\it Tycho} photometry (\blue{\citealt{hog2000}}) and solar analogs. The inside panel shows a comparison between gyro and seismic--ages for the {seismic sample} stars with a shaded region showing a $\pm~20\%~$ error.}

\caption{Angular velocity of dwarfs with 1.0 \msun~as a function of stellar age. Filled circles represent the seismic sample dwarfs. Open circles represent solar twin candidates. The shaded region represents the gyro relations (\blue{\citealt{Barnes07}}) with $(B-V) = 0.642$ for the Sun, with $0.95~\leq M/M_{\odot} \leq~1.10$. Diamonds indicate 16~Cyg~A and B.}

131 — 1407.2324

\caption{\textbf{Magneto-mechanically mediated quantum state transfer between a microwave qubit and light.} The main plot shows the Rabi oscillation between an optical resonator (OR) and the flux qubit (SQ) with small excitation of the mechanical resonator (MR). In addition, we show the states of the various sub-systems at the times $\kappa t=0.0, 0.39, 0.85$, showing the Wigner functions for the OR \& MR and the real part of the reduced density matrix of the SQ (imaginary components are zero). The system is initially prepared in the state$\rho_i=\rho^{SQ,OR}_i\otimes\rho^{MR}_i$, with $\rho^{SQ,OR}_i$ a pure excited(vacuum) state for the qubit(optical) modes, $|\psi_{Q/O}\rangle_i=|e,0\rangle$, and $\rho^{MR}_i$ is a mechanical thermal state with $\bar{n}=0.2$ phonons. {\color{black} We assume no cooling here and the initial phonon occupation is $n_b=0.2$, with $\gamma_q/2\pi=0.02$ \mega\hertz~ and $\kappa/2\pi=10$ \mega\hertz.} Other parameters are $\Delta_c=\Omega=0.9\omega_m =33\kappa$ and $\Lambda=G=5\kappa$. The photonic and phonon spaces are truncated to $N_a=4$ and $N_b=6$, respectively. }

132 — 1407.2564

\caption{(Color Online) ($a$)~Plot of the averaged directed current for a non-interacting system driven by a periodic potential with periodicity $1\s$, for densities $\rho =0.1$ ({\Large $\circ$}), $0.2$ ({\large $\square$}), $0.3$ ($\bdiamond$) and $0.6$ ($\triangle$). The resonance in the measured flux occurs at a fixed frequency $f_0 = 3.5 $. %% ($b$)~Plot of the measured flux for an interacting system driven by a periodic potential with peroidicity $1\s$, for densities $\rho =0.1$ ({\color{myblue} \large $\circ$}), $0.2$ ({\color{mypurple} $\square$}), $0.3$ ({\color{myokker} $\bdiamond$}) and $0.6$ ({\color{myblue} $\triangle$}). The solid lines are fit to the data using the functional form of Eq.~(2) in the main text. %% ($c$)~Plot of the measured flux for an interacting system driven by a periodic potential with peroidicity $1\s$, for densities $\rho =0.8$ ({\color{myblue} \Large $\bullet$}), $0.92$ ({\color{mypurple} $\blacksquare$}) and $0.98$ ({\color{myokker} $\fdiamond$}). The solid lines are fit to the data using the functional form of Eq.~(2) in the main text. The inset depicts the dependence of the resonance frequency on the density of the suspension. The solid line is a linear fit of $f_0 \sim (1-\rho/\rho_c)$ to the data with $\rho_c\approx 1.13$. %% ($d$)~Plot of the measured order parameter for the reciprocal lattice vectors $\mathbf{G}_2$ (empty symbols) and $\mathbf{G}_1$ (inset, filled symbols) for densities $\rho = 0.94$~({\color{myblue} \Large $\bullet$, $\circ$}) and $0.96$~({\color{mypurple}\footnotesize $\blacksquare$,$\square$}).}

\caption{Plot of the measured diffusion coefficient,normalized by the diffusion coefficient of a free system $D_0(\rho)$, for an interacting system driven by a periodic and a commensurate potenial, along (in ($a$) and ($b$)) and perpendicular (in ($c$) and ($d$)) to the direction of the external drive , as a function of frequency for densities of the suspension $\rho =0.2$ ({\color{myblue}{\Large $\circ$}}), $0.5$ ({\color{mypurple} $\square$}), $0.8$ ({\color{myokker}{$\bdiamond$}}), $0.92$ ({\color{myblue}{\Large $\bullet$}}), $0.94$ ({\color{mypurple} $\blacksquare$}) and $0.96$ ({\color{myokker}{$\fdiamond$}}). The solid lines are guide to the eye. }

133 — 1407.2854

\caption{\label{fig:measure comp}Modularity and Compartmentalization coefficient values for graphs simulated from the generative model with $N = 100$, $D_M = 0.502$ and only one group with more than one member with values averaged over 1000 simulations for $\rho = 0$ (black), \color{red} $\rho = 0.25$ (red), \color{green} $\rho = 0.5$ (green), \color{blue} $\rho = 0.75$ (blue), \color{ProcessBlue} $\rho = 1$ (light blue).}

134 — 1407.3216

\caption{ \label{fig:reflecting_chi0_0.0} (a) Schematic diagram of the model system. A sphere of radius $R$ has a surface which is partially catalytic (black) and partially inert (grey). The extent of the catalytic cap is parametrized by $\chi_{0} \equiv -\cos(\psi)$. %Over the catalytic part, a chemical product (green discs) is produced and diffuses in the %solution. Green discs \textcolor{black}{indicate} a diffusing chemical product. The system is bounded by an inert wall. The height $h$ and the orientation angle $\theta$ specify the configuration of the sphere. %The configuration of the Janus sphere is specified by the height $h$ of its center %above the wall and by the orientation angle $\theta$ of the symmetry axis %f the Janus sphere relative to the wall normal. (b) Phase plane for half coverage by catalyst ($\chi_{0} = 0$) of the Janus particle, where the color encodes $U_{y}/U_{0}$. %, i.e., the component of the normalized velocity of the particle parallel to the wall. (c) A typical trajectory of such a Janus particle. % Upon moving away from the wall, % *** change *** %(right side of the panel) %the orientation of the particle attains a value $\theta %\lesssim \pi/2$. The initial configuration % of the particle is $h_{0}/R = 5$ and $\theta_{0} = 120^{\circ}$, which is indicated in (b) with the symbol \textcolor{red}{$\boxtimes$}. }

\caption{\label{fig:reflecting_chi0_0.4} (a) Phase plane for $\chi_0 = 0.4$. Colors indicate $U_{y}/U_{0}$. There is a dynamical attractor at $h_{eq}/R = 1.42$ and $\theta_{eq} = 119.9^{\circ}$ (solid red circle). % the particle moves along the wall with constant $h$ and $\theta$. The dashed curve shows the analytically estimated ``slow manifold'' $U_{z} = 0$ (see main text) with a numerically fitted prefactor. %In the regions of the phase space $\theta < \pi/4$ and %$\theta > 4 \pi/5$ not shown here, the particle either escapes from or ``crashes'' into %the wall. (b) Flow field in the laboratory frame (white streamlines) and the solute number density $c/c_{0}$ (colors) associated with the sliding state. (c) A typical trajectory with the initial configuration $h_{0}/R = 6$ and $\theta_{0} = 3 \pi/4$ in the basin of attraction for sliding (\textcolor{red}{$\boxtimes$} in (a)). (d) A trajectory with $h_{0}/R = 1.2$ and $\theta_{0} = 107.5^{\circ}$ (\textcolor{red}{$\circledast$} in (a)). (e) Variation of the locations of the attractor (\textcolor{red}{$\bullet$}) and of the saddle point (\textcolor{blue}{$\circ$}) with $\chi_{0}$.}

\caption{\label{fig:reflecting_chi0_0.9} (a) Phase plane for $\chi_0 = 0.9$. Colors indicate $U_{y}/U_{0}$. There is an attractor at $h_{eq}/R = 1.78$ and $\theta_{eq} = 180^{\circ}$ (solid red circle). This is a ``hovering'' state in which the particle remains fixed in space. The dashed curve shows the fitted slow manifold. %In the section $\theta < \pi/2$ of the phase space %(not shown) the particle escapes. (b) Flow field (white streamlines) and solute number density $c/c_{0}$ (colors) for the hovering state. These fields are rotationally symmetric about the axis $x = 0$, $y = 0$. (c) An exemplary trajectory with initial configuration $h_{0}/R = 6$ and $\theta_{0} = 3\pi/4$ \textcolor{red}{$\boxtimes$} in (a)). % *** change *** % which belongs to the basin of attraction for hovering. (d) The trajectory for $h_{0}/R = 1.2$ and $\theta_{0} = 130^{\circ}$ (\textcolor{red}{$\circledast$} in (a)). (e) Variation of the height $h_{eq}/R$ of the hovering state with catalyst coverage $\chi_{0}$. %The numerical data broadly agree with an analytical estimate obtained by %modeling the effect of the wall in terms of an image source monopole. }

\caption{\label{fig:chi0_0.0_inert_0.9} (a) Phase plane for half coverage ($\chi_{0} = 0$) and unequal surface mobilities: $\beta = b_{inert}/b_{cap} = 0.9$. There is an attractor at $h_{eq}/R = 1.85$ and $\theta_{eq} = 98.2^{\circ}$ ({\color{red}$\bullet$}). Colors indicate $U_{y}/U_{0}$, where $U_{0} \equiv |b_{cap} \kappa/D|$. (b) Variation of the attractor location with $\beta$ for half coverage. (c) Trajectory with initial configuration $h_{0}/R = 1.2$ and $\theta_{0} = \pi/2$ (\textcolor{red}{$\circledast$} in (a)). % *** change **** %The particle initially moves away from the wall before rotating its cap away from the wall and %approaching a steady height and orientation. (d) Trajectory with $h_{0}/R = 6$ and $\theta_{0} = 125^{\circ}$ (\textcolor{red}{$\boxtimes$} in (a)). This is the same initial configuration as in Fig. \ref{fig:reflecting_chi0_0.0}(c).}

135 — 1407.3779

\caption{QMC results for the weigths $p_m$ computed for XXZ chains with anisotropy $\Delta=-0.9$. ({\color{red}{$\bullet$}}) $L=256$, PBC and $T=0$ with a bipartition at $\ell=128$. ($\circ$) $\ell=128$, OBC, and $T=T_{\rm ent.}$ Eq.~\eqref{eq:Teff}.}

136 — 1407.4083

\caption{\label{fig:MasterPlot}Phase space of convergence to quantum mechanics for $H=2\sigma_{z}$, $\rho\left(0\right)=\left\{ \left\{ 0.16,0.08,0.06\right\} ,\left\{ 0.23,0.3,0.17\right\} \right\}$, $\phi\left(0\right)=\left\{\left\{0,\Delta\phi_{0},2\Delta\phi_{0}\right\},\left\{\frac{\pi}{2}+\Delta\phi_{0},\frac{\pi}{2},\frac{\pi}{2}+\frac{\Delta\phi_{0}}{2}\right\}\right\}$, and $F$ given by Equation \ref{eq:spikeF} with varying $c$ {\red (characterizing the compactness of the F-kernel)} and $\Delta\phi_{0}$ {\red (the initial phase difference per value of the observable)}. Blue circles indicates convergence, red squares indicates lack of convergence, and purple diamonds means spin up converges, but spin down does not.}

137 — 1407.5266

\caption{Uniform mesh \blue{with} $\Delta x=\Delta y$ and the \blue{finite volume} cell constructed on it (filled with gray). \blue{The filled circles are lattice Boltzmann nodes. The time evolution of the population densities at the central node is controlled by the projected (onto the surface normal) sum of all the fluxes entering/leaving the gray region.}}

138 — 1407.5536

\caption{\textbf{(a):} Pulse sequence diagram for implementation of noiselet encoding in 2D imaging, where $G_{ss}$ is the gradient in slice (z) direction, $G_{pe}$ is the gradient in phase encoding (y) direction, and $G_{ro}$ is the gradient in readout (x) direction. The RF pulse duration is 5.12 ms and the flip angle is 10$^\circ$, which excites a noiselet profile along y-direction. The 180$^\circ$ refocusing pulse is used to select the desired slice in z-direction. A new RF pulse is executed for every new TR and the complete acquisition of all noiselet basis functions requires 256 different RF pulses derived from the noiselet measurement matrix. \textcolor[rgb]{0.00,0.00,0.00}{\textbf{(b):} Pulse sequence diagram for implementation of noiselet encoding in 3D imaging. The RF pulse excites a noiselet profile along y-direction and gradient blips are used along slice encode direction to encode slice direction with Fourier bases.}}

\caption{\textcolor[rgb]{0.00,0.00,0.00}{The mean relative error and standard deviation} (vertical bar) versus the number of receive channels for acceleration factors of 2 and 3, showing that the error increases as the number of channels decreases. Noiselet encoding outperforms Fourier encoding for both acceleration factors when the number of channels is more than two. However for a single channel, noiselet encoding outperforms Fourier encoding only for the acceleration factor of 2.}

\caption{\textcolor[rgb]{0.00,0.00,0.00}{The mean relative error and standard deviation }(vertical bar) versus the acceleration factor in MCS-MRI highlighting that noiselet encoding consistently outperforms Fourier encoding.}

\caption{\textbf{(a), (b) and (c):} are the plots of the mean relative error as a function of the signal to noise ratio (SNR) for different number of measurements. When the SNR is greater than 20 dB, the noiselet encoding outperforms Fourier encoding in the presence of noise for all acceleration factors; \textcolor[rgb]{0.00,0.00,0.00}{\textbf{(d):} show the brain images with SNR of 10, 20, 30 and 50 dB.}}

\caption{\textcolor[rgb]{0.00,0.00,0.00}{\textbf{(a)} and \textbf{(b):} Two slices of image reconstructed using 3D GRE noiselet encoding sequence, the up/down direction is noiselet encoding direction and left/right direction if Fourier encoding direction; \textbf{(c)} and \textbf{(d):} two slices of image reconstructed with Fourier encoding using 3D GRE sequence.}}

139 — 1407.5864

\caption{Magnetization along the applied field of an Fe$_2$P crystal. (a) Magnetization as function of field applied along the $c$-axis (solid curves) and perpendicular to the $c$-axis (dashed curves) for three different temperatures; $188\,\mathrm{K}$, $216\,\mathrm{K} \approx T_{\mathrm C}$ and $276\,\mathrm{K}$. (b) Temperature dependence of the magnetization for fields along (filled symbols) and perpendicular to (open symbols) the $c$-axis (\textcolor{red}{$\circ$} = 0.01 T, \textcolor{green}{$\Box$} = 0.1 T, and \textcolor{blue}{$\rhd$} = 1 T). The inset shows the low-field curves for $\bf{H}\, \perp\, \bf{c}$. $T_\mathrm{C} (0\,{\rm T})$ is indicated by vertical dashed lines.}

140 — 1407.7510

\caption{ (Color online) Swapping protocol to compensate the momentum displacement shown in Fig. 2(b) and resulting gate performance. (a) Photons in the interacting rails $\left(1_{C},\,1_{T}\right)$ are stored as collective excitations and excited to the Rydberg levels $|r\rangle$ as described in Fig.~\ref{Scheme}. They are brought back to the spin state $|s\rangle$ after half of the interaction time ($\frac{t}{2}$). (b) Tilted control fields swap the relative positions of the two collective excitations using non-Rydberg EIT. (c) The collective excitations are re-excited to the Rydberg levels, interact for $\frac{t}{2}$, and are de-excited again. The photons are retrieved using non Rydberg EIT. (d) Gate fidelity as a function of the separation between the collective excitations. Solid and hollow circles are with and without the swapping protocol respectively. The spatial shape of the collective excitations is the same as in Fig. 2. (e) Gate efficiency (circles) and interaction time required for creating a $\pi$ phase shift (squares) as a function of the separation. The efficiency does not include photon storage and retrieval, see text. One sees that increasing the separation yields higher fidelity, but lower efficiency, because the weaker interaction for greater separations requires longer interaction times and hence more loss due to thermal motion and the finite lifetime of the Rydberg states. Using the swapping protocol, both high fidelity and high efficiency can be achieved.}\label{Swapping} \end{figure} We analyze the expected gate performance using the concepts of (conditional) fidelity and efficiency. Analogous concepts are commonly used in the context of quantum storage \cite{SimonEPJD}. The conditional fidelity quantifies the performance of the gate, conditioned on successful retrieval of both photons. The effects of photon loss are discussed in terms of efficiency below. Following the treatment in \cite{AdamStored}, the conditional fidelity of a gate operating on the initial state $\left|\phi\right\rangle=\frac{1}{2}\left(\left|0_{C}\right\rangle+\left|1_{C}\right\rangle\right)\left(\left|0_{T}\right\rangle+\left|1_{T}\right\rangle\right)$ can be quantified as $F=\sqrt{\left\langle \phi'\right|\rho\left|\phi'\right\rangle }$. This definition characterizes the gate's outcome $\rho$, relative to the ideal output $\left|\phi'\right\rangle =(\left|00\right\rangle +\left|01\right\rangle +\left|10\right\rangle -\left|11\right\rangle )/2$. Since the many-body interaction only affects the interacting pair, the fidelity can be rewritten as $F=\sqrt{(9-3(\zeta+\zeta^{*})+\left|\zeta\right|^{2})/16}$ where $ \zeta=\left\langle \Psi_{t_{0}}\right|e^{-i\hat{H}_{int}t}\left|\Psi_{t_{0}}\right\rangle$ with $|\Psi_{t_0}\rangle$ as given in Eq. (1) and $\hat{H}_{int}$ as defined above. It is clear from Fig. 2(b) that the momentum displacement and the entanglement-related profile deformation will affect the value of $\zeta$ and hence of $F$. Controlling these effects is essential for achieving high gate fidelity. \begin{figure} \scalebox{0.33}{\includegraphics*{entanglement5.pdf}} \caption{(Color online) Effects of unwanted entanglement on gate fidelity. (a) The fidelity has a non-isotropic dependence on the width of the collective excitations. Here the collective excitations are separated by 21~$\mu$m and their spatial profile has the same initial width of 8~$\mu$m in all directions. Compressing the width parallel to the separation ($w_{\shortparallel}$) has a significant impact on the fidelity (circles). In contrast, compressing the width perpendicular to the separation ($w_{\perp}$) has a negligible effect (triangles). (b) The fidelity reduction $1-F$ is proportional to the entanglement, quantified by the Von Neumann entropy. Here the momentum displacement is compensated by the swapping protocol of Fig. 3, leaving the unwanted entanglement as the main source of infidelity.}\label{Entanglement} \end{figure} We propose a swapping protocol to compensate the destructive effects of momentum displacement, see Fig. 3. Since the momentum displacement $\boldsymbol{k}_{D}\propto\Delta\boldsymbol{x}_{0}t$ is proportional to the separation vector, swapping the relative position of the collective excitations $\left(\Delta\boldsymbol{x}_{0}\rightarrow-\Delta\boldsymbol{x}_{0}\right)$ in the middle of the interaction time reverses the rate of momentum displacement creation. The compensation of the momentum displacement after swapping can be seen in Fig. 2(c), and its beneficial effect on the gate fidelity in Fig. 3(d). The swapping protocol is relatively robust to positioning errors. In an example where the collective excitations are separated by 21~$\mu$m, an averaged Gaussian error of 1~$\mu$m in the parallel dimension reduces the average fidelity by 1\%, see also the supplemental materials \cite{supp}. Errors in the perpendicular dimension are much less critical \cite{supp}. It is important to also consider photon loss. Photon loss that is uniform over the four rails has no effect on the conditional fidelity as defined above. It can therefore be discussed independently in terms of the efficiency $\eta$, which is the probability of retrieving each photon after the gate operation. Non-uniform loss terms in our scheme can be made uniform by adding external sources of loss to certain rails, see supplementary information \cite{supp}. Two important sources of loss are atomic thermal motion \cite{supp,Thermal-motion} and the finite life time of the Rydberg levels (1.15 ms for $|102\, S_{1/2}\rangle$ and 1.18 ms for $|103\, S_{1/2}\rangle$) \cite{Life-time}. Their effects on the efficiency are shown in Fig.~\ref{Swapping}(e) for different interaction times in an ensemble cooled to T=0.1~$\mu$K. Considering the separation of interacting rails, there is a trade-off between fidelity and efficiency. A small separation improves the efficiency by reducing the interaction time (see Fig.~\ref{Swapping}(e)), but the resulting stronger interaction causes more entanglement and momentum displacement, which reduces the fidelity (see Fig.~\ref{Swapping}(d)). The swapping protocol makes it possible to achieve high fidelity and high efficiency simultaneously. Another significant source of inefficiency comes from the process of storage and retrieval of single photons. A conservative estimate of the associated efficiency for the whole protocol (including the swapping) can be obtained by applying the photon's storage and retrieval efficiency twice \cite{Efficiency}. This corresponds to the use of two separate MOTs for storing photons before and after swapping. Based on this estimate the overall efficiency for a density of $\rho=4\times10^{12}$~cm$^{-3}$ (corresponding to an optical depth $d \approx 100$) is about 70\% . Increasing the density to $\rho=3.8\times 10^{13}$~cm$^{-3}$ ($d \approx 750$) improves the efficiency of repeated storage and retrieval to 95\%. In practice using a single MOT is likely to both be more practical and lead to higher efficiency than these estimates because in this case the stationary excitations only have to be converted into moving excitations (but not all the way into photons) at the intermediate stage. We have shown how to compensate the effect of momentum displacement on the fidelity. The other destructive effect of the interaction that reduces the fidelity is the creation of unwanted entanglement between the collective excitations. Entanglement reduces the fidelity by deforming the two-excitation wave function in momentum space, see Fig.~\ref{SiKy}(b) (see also Fig.~5(b) in the supplemental materials \cite{supp}). Comparing the eccentricities of the ellipses in parallel and perpendicular direction obtained from Eq.~(\ref{SiPP}), $\frac{e_{\shortparallel}^{2}}{e_{\perp}^{2}}\backsim\frac{49w_{\shortparallel}^{4}}{w_{\perp}^{4}}$, one sees that the deformation is much stronger for the parallel dimension. Therefore, compression of the collective excitations parallel to their separation can reduce the unwanted entanglement while leaving room for extra atoms in the perpendicular dimensions in order to preserve the directionality of the collective emission \cite{directionality1,directionality}. The highly non-isotropic effects of profile compression on the fidelity are shown in Fig.~\ref{Entanglement}(a). The achievable width compression is mainly limited by diffraction. In order to show the relation between fidelity and entanglement even more clearly we calculate the Von Neumann entropy of the output state. Fig.~\ref{Entanglement}(b) shows that fidelity reduction and entanglement are indeed proportional. In conclusion, we have proposed a photon-photon gate protocol that uses stationary collective Rydberg excitations, but does not rely on photon blockade. We have shown that unwanted effects due to the distance-dependence of the interaction are important but can be overcome, making it realistic to achieve a gate operation with high fidelity and efficiency. {\it Acknowledgments.} We acknowledge financial support from AITF and NSERC. We thank B. He, A. Kuzmich, D. Paredes-Barato and B. Sanders for fruitful discussions. \begin{thebibliography}{10} \bibitem{PhotonGate-1}H. Schmidt and A. Imamoglu, Optics Letters \textbf{21}, 23 (1996) \bibitem{PhotonGate-2}M.D. Lukin and A. Imamoglu, Phys. Rev. Lett \textbf{84}, 7 (1999) \bibitem{PhotonGate-5}Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, and H. J. Kimble, Phys. Rev. Lett. \textbf{75}, 4710 (1995) \bibitem{PhotonGate-6}L.-M. Duan and H. J. Kimble, Phys. Rev. Lett. \textbf{92}, 127902 (2004) \bibitem{PhotonGate-4} A. Rispe, B. He, and C. Simon, Phys. Rev. Lett. \textbf{107}, 043601 (2011) \bibitem{NutralAtomGateProposal-11}M. Saffman and T. G. Walker, Phys. Rev. A \textbf{72}, 022347 (2005) %interaction \bibitem{NutralAtomGateProposal-1} D. Jaksch, J.I. Cirac, P. Zoller, S.L. Rolston, R. Cote, and M. D. Lukin, Phys. Rev. Lett. \textbf{85}, 2208 (2000) %Blockade %interaction \bibitem{NutralAtomGateProposal-2}M. D. Lukin, M. Fleischhauer, R. Cote, L. M. Duan, D. Jaksch, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. \textbf{87}, 037901 (2001) %Blockade \bibitem{NutralAtomGateProposal-3} M. M\"uller, I. Lesanovsky, H. Weimer, H. P. B\"uchler, and P. Zoller, Phys. Rev. Lett.\textbf{102}, 170502 (2009) %Blockade \bibitem{NutralAtomGateProposal-4}L. Isenhower, M. Saffman, and K. K. M\o lmer, Quant. Info. Pro. \textbf{10}, 755 (2011) %Blockade \bibitem{NutralAtomGateImplement-1}M. Saffman, T. G. Walker, and K. M\o lmer, Rev. Mod. Phys. \textbf{82}, 2313 (2010). %Blockade \bibitem{NutralAtomGateImplement-2}L. Isenhower, E. Urban, X. L. Zhang, A. T. Gill, T. Henage, T. A. Johnson, T. G. Walker, and M. Saffman, Phys. Rev. Lett. \textbf{104}, 010503 (2010) %Blockade \bibitem{NutralAtomGateImplement-3}T. Wilk, A. Ga\"etan, C. Evellin, J. Wolters, Y. Miroshnychenko, P. Grangier, and A. Browaeys, Phys. Rev. Lett.\textbf{104}, 010502 (2010) %Blockade \bibitem{PhotAtomIntera-1}Y. O. Dudin and A. Kuzmich, Science \textbf{336}, 887 (2012), \bibitem{PhotAtomIntera-2}T. Peyronel, O. Firstenberg, Q. Y. Liang, S. Hofferberth, A. V. Gorshkov, T. Pohl, M. D. Lukin, and V. Vuleti\'{c}, Nature \textbf{488}, 57 (2012) \bibitem{PhotAtomIntera-3}D. Maxwell, D. J. Szwer, D. Paredes-Barato, H. Busche, J. D. Pritchard, A. Gauguet, K. J. Weatherill, M. P. A. Jones, and C. S. Adams, Phys. Rev. Lett. \textbf{110}, 103001 (2013). \bibitem{PhotAtomIntera-4}S. Sevin\c cli, C. Ates, T. Pohl, H. Schempp, C. S. Hofmann, G. G\"unter, T. Amthor, M. Weidem\"uller, J. D. Pritchard, D. Maxwell, A. Gauguet, K. J. Weatherill, M. P. A. Jones, and C. S. Adams, At. Mol. Opt. Phys.\textbf{44}, 184018 (2011) \bibitem{PhotAtomIntera-5}V. Parigi, E. Bimbard, J. Stanojevic, A.J. Hilliard, F. Nogrette, R. Tualle-Brouri, A. Ourjoumtsev, and P. Grangier, Phys. Rev. Lett. \textbf{109}, 233602 (2012) \bibitem{PhotAtomIntera-6}C. S. Hofmann, G. G\"unter, H. Schempp, M. Robert-de-Saint-Vincent, M. G\"arttner, J. Evers, S. Whitlock, and M. Weidem\"uller, Phys. Rev. Lett.\textbf{110}, 203601 (2013). \bibitem{many-body}Y. O. Dudin, L. Li, F. Bariani, and A. Kuzmich, Nature Physics \textbf{8}, 790 (2012) \bibitem{PhotBlock-1}A. V. Gorshkov, J. Otterbach, M. Fleischhauer, T. Pohl, and M. D. Lukin, Phys. Rev. Lett. \textbf{107}, 133602 (2011). \bibitem{PhotBlock-2}O. Firstenberg, T. Peyronel, Q. Y. Liang, A. V. Gorshkov, M. D. Lukin, and V.Vuleti\'{c}, Nature \textbf{502}, 71 (2013) \bibitem{Phot-Int1}I. Friedler, D. Petrosyan, M. Fleischhauer and G. Kurizki, Phys. Rev. A \textbf{72}, 043803 (2005) \bibitem{Phot-Int2}E. Shahmoon, G. Kurizki, M. Fleischhauer and D. Petrosyan, Phys. Rev. A \textbf{83}, 033806 (2011) \bibitem{Phot-Int3} B. He, A. MacRae, Y. Han, A. I. Lvovsky, and C. Simon, Phys. Rev. A \textbf{83}, 022312 (2011) \bibitem{Phot-int4} B. He, A. V. Sharypov, J. Sheng, C. Simon, and M. Xiao, Phys. Rev. Lett. \textbf{112}, 133606 (2014) % %\bibitem{Barianidephasing}F. Bariani, Y. O. Dudin, T. A. B. Kennedy, and A. Kuzmich, Phys. Rev. Lett. \textbf{108}, 030501 (2012) \bibitem{AdamStored}D. Paredes-Barato and C. S. Adams, Phys. Rev. Lett. \textbf{112}, 040501 (2014) %\bibitem{Molecule} V. Bendkowsky, B. Butscher, J. Nipper, J. P. Shaffer, %R. Low, T. Pfau, Nature 458, 1005 (2009) %\bibitem{Angular}A. Reinhard, T. Cubel Liebisch, B. Knuffman, and %G. Raithel, Phys. Rev. A 75, 032712 (2007) \bibitem{ChuangYamamoto}I. L. Chuang and Y. Yamamoto, Phys. Rev. A \textbf{52}, 3489 (1995) \bibitem{supp} The supplemental materials are appended to this letter. \bibitem{Rydberg-level-1}J. Han and T. F. Gallagher, Phys. Rev. A \textbf{79}, 053409 (2009). \bibitem{Rydberg-level-2} T. G. Walker and M. Saffman, Phys. Rev. A \textbf{77}, 032723 (2008) \bibitem{Sign of interaction} A. Reinhard, T. C. Liebisch, B. Knuffman, and G. Raithel, Phys. Rev. A \textbf{75} 032712 (2007) \bibitem{Sign of interaction-2} F. Maucher, N. Henkel, M. Saffman, W. Krolikowski, S. Skupin, and T. Pohl, Phys. Rev. Lett. \textbf{106} 170401(2011) \bibitem{SimonEPJD} C. Simon {\it et al.}, Eur. Phys. J. D {\bf 58}, 1 (2010). \bibitem{Thermal-motion}S. D. Jenkins, T. Zhang, and T. A. B. Kennedy, J. Phys. B \textbf{45}, 124005 (2012) \bibitem{Life-time}T.F. Gallagher, {\it Rydberg Atoms} (Cambridge University Press, Cambridge, 2005) \bibitem{Efficiency} A.V. Gorshkov, A. Andr\'{e}, M.D. Lukin, and A. S. S\o rensen, Phys. Rev. A \textbf{76}, 033805 (2007). \bibitem{directionality1}M. Saffman and T. G. Walker, Phys. Rev. A \textbf{66}, 065403 (2002) \bibitem{directionality}Y. Han, B. He, K. Heshami, C. Li, and C. Simon, Phys. Rev. A \textbf{81}, 052311 (2010) %\bibitem{DensityMatrix} B. He and A. Scherer, Phys. Rev. A \textbf{85} 033814 (2012) \end{thebibliography} \pagebreak \widetext %\begin{widetext} \begin{center} \textbf{\large Supplemental Materials for ``Photon-photon gate via the interaction between two collective Rydberg excitations''} \end{center} \setcounter{table}{0} \setcounter{equation}{0} \setcounter{page}{1} \makeatletter \renewcommand{\bibnumfmt}[1]{[SI#1]} \renewcommand{\citenumfont}[1]{SI#1} \title{Supplemental Materials for ``Photon-photon gate via the interaction between two collective Rydberg excitations''} \author{Mohammadsadegh Khazali, Khabat Heshami and Christoph Simon} \affiliation{Institute for Quantum Science and Technology and Department of Physics and Astronomy, University of Calgary, Calgary T2N 1N4, Alberta, Canada} \date{\today} \maketitle \section{Effects of Interaction on the wave function in the perpendicular dimensions} Fig. 5 shows the numerical evaluation of $\left|\psi_{k_{1\perp},k_{2\perp}}\right|^{2}$, the modulus squared of the two-excitation wave function in the dimensions perpendicular to the separation. While the second order of the interaction changes the Gaussian cross section from circular to elliptical by entangling the two excitations, the first order does not have any effect in this dimension. \begin{figure}[h]\label{SiPerpendicular} \scalebox{0.5}{\includegraphics*{perpendicular.pdf}} \caption{Numerical results for $\left|\psi_{k_{1\perp},k_{2\perp}}\right|^{2}$. There is no momentum displacement in this dimension. The parameters are the same as for Fig. 2.} \end{figure} \section{Sensitivity of Fidelity to positioning errors in Swapping} Fig. 6 shows that the swapping protocol is more sensitive to positioning errors for the collective excitations in the parallel dimension than in the perpendicular dimension. \begin{figure}\label{errorinswapping} \scalebox{0.35}{\includegraphics*{errorinswapping.pdf}} \caption{Average fidelity reduction as a function of positioning error for the swapping protocol in (a) parallel and (b) perpendicular dimension. The parameters are the same as for Fig.~2.} \end{figure} \section{``Frozen Collision''} Fig. 7 shows the redistribution of the momentum vectors of the collective excitations due to the interaction. One sees that collective excitations that are created by the storage of co-propagating photons will yield diverging photons upon retrieval. \begin{figure}[h!]\label{AngularDestribution} \scalebox{0.3}{\includegraphics{AngularDistribution.pdf}} \caption{Angular distribution of the momenta of the collective excitations before (black for both) and after (red and blue) interaction. The parameters are the same as for Fig. 2.} \end{figure} \section{Gate Efficiency: thermal motion and uniform loss} \subsection{Thermal motion} The efficiency factor due to the thermal motion of the atoms is given by $\eta_{_{th}}=\frac{1}{(1+(\frac{t}{\xi})^{2})^{2}}exp[\frac{-t^{2}/\tau^{2}}{(1+(t/\xi)^{2})}]$ \cite{key-1}, where $\tau=\frac{\Lambda}{2\pi v}$ is the dephasing time scale, which is determined by the wave length $\Lambda$ of the collective excitations and the thermal speed $v$, and $\xi=\frac{w}{v}$ is the time scale on which an atom traverses the width $w$ of a collective excitation. \subsection{Uniform loss} Photon loss that is uniform over the four rails has no effect on the conditional fidelity and can be quantified independently in terms of the efficiency. Since only the interacting rails ($\left|1_{C}\right\rangle ,\left|1_{T}\right\rangle $) are excited to Rydberg levels, they experience an extra loss due to the life time of the Rydberg level. Furthermore, the shorter wavelength of the Rydberg excitations in the interacting rails creates a stronger loss due to the atomic thermal motion compared to the non-interacting rails. Finally, the extra process of storage and retrieval during the swapping of the interacting rails causes an extra loss of efficiency in these rails. The loss can be made uniform by adding a controlled external source of loss on the non-interacting rails ($\left|0_{C}\right\rangle ,\left|0_{T}\right\rangle $). \begin{thebibliography}{1} \bibitem{key-1}S. D. Jenkins, T. Zhang, T. A. B. Kennedy, J. Phys. B 45, 124005 (2012) \end{thebibliography} \end{document} }

141 — 1407.7516

\caption{\footnotesize Left: power spectrum of \hat{} (black) with the fitted model over-plotted (red; cf. \eqref{eq:limitspec}). Right: power spectrum of \hat{}\(black) over-plotted with the optimum fit to the background (red; cf. \eqref{eq:bg}). The light-grey part up to $\rm 100\, \mu Hz$ was not included in the fit. The fit includes, besides the Gaussian envelope from $p$-modes centred around $\nu_{\rm max}\approx1115\,\rm \mu Hz$, a granulation component (\textcolor[rgb]{0, 0.5, 0}{\textbf{P}$\rm _G$}; green) and a white/shot noise (\textcolor{black}{\textbf{P$\rm _S$}}; black) level. The dashed red line shows the background fit without the Gaussian envelope.}

\caption{\footnotesize Top: autocorrelation function (ACF) of the \hat{}\time series (black). The dashed horizontal line gives the expectation value for random independent and identically distributed values, while the grey curves give the large-lag$95\%$ confidence levels. Bottom: low-frequency end of the power spectrum in units of period.}

142 — 1408.0343

\caption{Facilitation in translational and orientational degrees of freedom. (A) Instanton time distribution $P_r(\Delta t)$ for translational excitations of size $a_r = 0.33l$ for $\phi =$ 0.68 ({\color{black} $\boldsymbol \blacksquare$}), $\phi =$ 0.73 ({\color{red} $\boldsymbol \bullet$}), $\phi =$ 0.76 ({\color{blue} $\boldsymbol \blacktriangle$}) and $\phi =$ 0.79 ({\color{green!50!black} $\boldsymbol \blacktriangledown$}). (B) Instanton time distribution $P_{\theta}(\Delta t)$ for rotational excitations of size $a_{\theta} = 10^{\circ}$ for $\phi =$ 0.68 ({\color{black} $\boldsymbol \square$}), $\phi =$ 0.73 ({\color{red} $\boldsymbol \circ$}), $\phi =$ 0.76 ({\color{blue} $\boldsymbol \triangle$}) and $\phi =$ 0.79 ({\color{green!50!black} $\boldsymbol \triangledown$}). (C) The function $\mu_{rr}(r,-t_r/2,t_r/2;a_r)$ for $a_r = 0.33l$ for various $\phi$s. The colors and symbols are identical to those in (A). (D) The function $\mu_{\theta\theta}(r,-t_{\theta}/2,t_{\theta}/2;a_{\theta})$ for $a_{\theta} = 10^{\circ}$ for various $\phi$s. The colors and symbols are identical to those in (B). (E) The translational displacement field $|\Delta \textbf{r}_i(t_r)|$ normalized by $a_r = 0.33l$.(F) The rotational displacement field $|\Delta \theta_i(t_{\theta})|$ normalized by $a_\theta = 10^{\circ}$. In (E) and (F), the top panels correspond to $\phi =$0.73 and the bottom panels correspond to $\phi =$0.79. (G) Facilitation volume for translational excitations, $v_{F}^{r}(t)$, for $a_r = 0.33l$ for various $\phi$s. The colors and symbols are identical to those in (A). (H) Facilitation volume for rotational excitations, $v_{F}^{\theta}(t)$, for $a_\theta = 10^{\circ}$ for various $\phi$s. The colors and symbols are identical to those in (B).}

\caption{Coupling between translational and rotational facilitation. (A) The off-diagonal functions $\mu_{r\theta}(r,-t_r/2,t_r/2;a_r)$ ({\color{red} $\boldsymbol \blacksquare$}) and $\mu_{\theta r}(r,-t_{\theta}/2,t_{\theta}/2;a_{\theta})$ ({\color{blue} $\boldsymbol \circ$}) for $\phi =$0.76. The values $a_r = 0.5l$ and $a_{\theta} = 15^{\circ}$ are chosen such that $\langle \Delta t\rangle_r \sim \langle \Delta t\rangle_{\theta}$. (B) Facilitation volume profiles for the off-diagonal functions in (C). (B) The asymmetry parameter $F(a_r,a_{\theta},t_m)$ for various combinations of $a_r$ and $a_{\theta}$. The corresponding values of $\langle \Delta t\rangle_r$ and $\langle \Delta t\rangle$ are enclosed in parentheses. (C) The maxima of the off-diagonal facilitation volumes, $v_{\text{max}}^{r}$ and $v_{\text{max}}^{\theta}$ as a function of $\phi$, normalized by their respective values at $\phi_0 =$ 0.68. The colors and symbols are identical to those in (A).}

\caption{Spatial decoupling of heterogeneities and prediction of reentrant glass transitions. (A) The coupling coefficient for facilitation, $C_{F}(a_r,a_{\theta},t_m)$, with $a_r = 0.5l$ and $a_{\theta} = 20^{\circ}$, as a function of $\phi$ for $\Delta u/k_BT =$ 0 ({\color{green!50!black} $\boldsymbol \blacksquare$}), $\Delta u/k_BT =$ 1.16 ({\color{red} $\boldsymbol \bullet$}) and $\Delta u/k_BT =$ 1.47 ({\color{blue!60!black} $\boldsymbol \blacktriangle$}). (B) The translational and rotational contributions to $C_{F}(a_r,a_{\theta},t_m)$, $C_{F}^{r}$ (filled blue symbols) and $C_{F}^{\theta}$ (hollow red symbols), respectively, as a function of $\phi$ for $\Delta u/k_BT =$ 0 (squares), $\Delta u/k_BT =$ 1.16 (circles) and $\Delta u/k_BT =$ 1.47 (triangles). (C) The coupling function for dynamical heterogeneities, $D(\Delta t)$, in the purely repulsive case, $\Delta u/k_BT =$ 0, for $\phi =$ 0.68 ({\color{black} $\boldsymbol \square$}), $\phi =$ 0.73 ({\color{red} $\boldsymbol \circ$}), $\phi =$ 0.76 ({\color{blue} $\boldsymbol \triangle$}) and $\phi =$ 0.79 ({\color{green!50!black} $\boldsymbol \triangledown$}). (D) The coupling coefficient for dynamical heterogeneities, $D_{\text{max}}$, as a function of $\phi$ for $\Delta u/k_BT =$ 0 ({\color{green!50!black} $\boldsymbol \blacksquare$}), $\Delta u/k_BT =$ 1.16 ({\color{red} $\boldsymbol \bullet$}) and $\Delta u/k_BT =$ 1.47 ({\color{blue!60!black} $\boldsymbol \blacktriangle$}). (E) The $\phi$ dependence of the concentration of translational excitations $c_r$ for $a_r = 0.5l$, for $\Delta u/k_BT =$ 0 ({\color{green!50!black} $\boldsymbol \blacksquare$}), $\Delta u/k_BT =$ 1.16 ({\color{red} $\boldsymbol \bullet$}) and $\Delta u/k_BT =$ 1.47 ({\color{blue!60!black} $\boldsymbol \blacktriangle$}). (F) The $\phi$ dependence of the concentration of rotational excitations $c_{\theta}$ for $a_{\theta} = 20^{\circ}$ for $\Delta u/k_BT =$ 0 ({\color{green!50!black} $\boldsymbol \square$}), $\Delta u/k_BT =$ 1.16 ({\color{red} $\boldsymbol \circ$}) and $\Delta u/k_BT =$ 1.47 ({\color{blue!60!black} $\boldsymbol \triangle$}). In (E) and (F), the curves are empirical fits of the form $\phi_0 + A(\phi_c - \phi)^{-1}$. (G) The translational glass transition $\phi_{c}^{r}$ ({\color{black} $\boldsymbol \blacksquare$}) and rotational glass transition $\phi_{c}^{\theta}$ ({\color{red} $\boldsymbol \circ$}) obtained from fits to the curves in (E) and (F), for various values of $\Delta u/k_{B}T$.}

143 — 1408.0604

\caption{The extended Bruggeman estimates of relative permittivity parameters of the HCM plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1} $} and spheroid orientation angle $\varphi \in \le 0, 90 \ri^\circ$. The refractive index $n_a = 1.1$, volume fraction $f_a = 0.3$, size parameter $ \eta = 0.3/\ko$, and eccentricity parameter $\rho = 3$. }

\caption{The angles $\alpha_{1,2}$ and $\beta_{1,2}$, along with the quantities $| W_{1,2} |$, plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1} $} for the spheroid orientation angles $\varphi = 60^\circ$ (red, solid curves), $30^\circ$ (green, dashed curves), and $90^\circ$ (blue, broken dashed curves). The refractive index $n_a = 1.1$, volume fraction $f_a = 0.3$, size parameter $\eta = 0.3/\ko$, and eccentricity parameter $\rho = 3$. }

\caption{The angles $\alpha_{1,2}$ and $\beta_{1,2}$ plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1}$ } for the eccentricity parameter values $\rho = 3$ (red, solid curves), $2$ (green, dashed curves), and $4$ (blue, broken dashed curves). The refractive index $n_a = 1.1$, volume fraction $f_a = 0.3$, size parameter $\eta = 0.3/\ko$, and spheroid orientation angle $\varphi = 60^\circ$. }

\caption{The angles $\alpha_{1,2}$ and $\beta_{1,2}$ plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1} $ } for the size parameter values $ \eta = 0.3/\ko$ (red, solid curves), $0.2/\ko$ (green, dashed curves), and $0.25/\ko$ (blue, broken dashed curves). The refractive index $n_a = 1.1$, volume fraction $f_a = 0.3$, spheroid orientation angle $\varphi = 60^\circ$, and eccentricity parameter $\rho = 3$. }

\caption{The angles $\alpha_{1,2}$ and $\beta_{1,2}$ plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1}$ } for the volume fractions $f_a = 0.3$ (red, solid curves), $0.35$ (green, dashed curves), and $0.4$ (blue, broken dashed curves). The refractive index $n_a = 1.1$, size parameter $\eta = 0.3/\ko$, spheroid orientation angle $\varphi = 60^\circ$, and eccentricity parameter $\rho = 3$. }

\caption{The angles $\alpha_{1,2}$ and $\beta_{1,2}$ plotted versus \blue{ $E^{dc}_3 \in \le -1, 1 \ri \times 10^8 $ $\mbox{V} \mbox{m}^{-1} $} for the refractive index values $n_a = 1.1$ (red, solid curves), $1.05$ (green, dashed curves), and $1.3$ (blue, broken dashed curves). The volume fraction $f_a = 0.3$, size parameter $\eta = 0.3/\ko$, eccentricity parameter $\rho = 3$, and spheroid orientation angle $\varphi = 60^\circ$. }

144 — 1408.1302

\caption{RO product conductivity (marked as {\color{red}@} in Fig.~\ref{figure:WS}) compared to pH ({\color{red}\$}) of dosed inlet water of the central FMU system, over several months. The vertical scale corresponds either to pH, or to conductivity in $\mu$S/cm.}

\caption{Product resistivity (marked as {\color{red}$\star$} in Fig.~\ref{figure:WS}) of each of the local PL systems (EH1, EH2, and EH3) from December 2011 to July 2012.}

\caption{EH2 pool resistivity (measured at the recirculation input to the PL, marked as {\color{red}\#} in Fig.~\ref{figure:WS}) and the average number of triggered photomultipliers per cosmic ray, during a period in October 2012, after repairs were made to the polishing loop. The correlation indicates an increasing attenuation length in the water with rise in resistivity.}

\caption{Resistivity of water at the PL input (marked as {\color{red}\#} in Fig.~\ref{figure:WS}) for EH1, EH2, and EH3 for the same time period as shown in Fig.~\ref{figure:ProductResist}. This is nominally the pool water resistivity, except for the spikes during short periods of filling.}

\caption{Dissolved oxygen content (in ppb O$_2$) measured at the inlet (marked as {\color{red}\&} in Fig.~\ref{figure:WS}) and outlet ({\color{red}$\sim$}) points of the EH1 water system.}

145 — 1408.1516

\caption{Multifrequency monitoring of the blazar 3C 454.3 in November 2010 ({\mttt MJD 55510 corresponds to November 10, 2010}). The panels show: (1) the $\gamma$-ray emission above 100 MeV as monitored by Fermi-LAT with a time bin of 3 hours. All the fluxes and spectra were obtained using the Fermi Science Tools, performing an unbinned likelihood analysis of the publicly available Fermi-LAT data set. The data analysis includes the Galactic and isotropic diffuse emission (using gal-2yearp7v6-v0.fits, iso-p7v6source), and all sources in the second LAT source catalog within $10^o$ from 3C454.3 (top panel); (2) X-ray emission in the range 2-10 keV as monitored by Swift and reported in \citealp{vercellone-1, wehrle} (second panel); (3) Mg II emission lines detected by \citealp{leon-tavares} (third panel); (4) optical GASP WEBT data of R-band (\citealp{vercellone-1}; fourth panel); (5) radio emission (230 GHz) monitoring reported in \citealp{vercellone-1, wehrle} (bottom panel). Note the 1-day optical flare around MJD 55510 with no detected counterpart in the X-ray and \gray bands.}

\caption{Multifrequency monitoring of the blazar 3C 454.3 during the exceptional \gray flaring in November 2010. Zoom of Fig. 2 showing the \gray emission above 100 MeV (top panel), and the R-band (bottom panel) during the peak $\gamma$-ray emission. Note around MJD 55517 the rise by a factor 4 in $\gamma$ rays, with the optical rising by a factor of only 2, whereas around MJD 55520 the flux in both bands rises by the same factor.}

\caption{Broad-band spectral energy distributions ({\vv SEDs}) computed for different states of 3C 454.3 {\mttt in November 2010} as indicated in the Appendix (Eqs. A3, A5, A7 and A9). The thick green curve gives the overall plasmoid radiation in the plateau state {\vv (MJD 55502-55516, Nov. 2-16, 2010). The green dotted curve shows the radiation of the dominant plasmoid with enhanced optical emission (MJD 55510, Nov. 10, 2010)}. This component adds to the plateau state to yield the optical flare SED {\vv (thin, green, solid line) that has no} \gray counterpart. The dotted magenta curve shows the mirror \gray IC radiation. This component summed with the plateau emission yields the total SED {\mttt on MJD=55517 (Nov. 17, 2010}, dashed magenta line). The thick red curve shows the peak flaring emission SED {\mttt (MJD 55520, Nov. 20, 2010)}. The blue dashed curve shows the post-flare SED after two days {\vv of mirror radiation decrease due to} cooling. Data in the \gray band are from {\mttt \textit{Fermi}-LAT; all other data are from } Vercellone et al. (2011). }

\caption{Model parameters for the November 2010 \gray flare of \source.}

146 — 1408.1803

\caption{(Color) Spectral tuning of photonic crystal cavity modes: (a) PL spectra taken before (black) and after (red) the fabrication of the M7-cavity (Fig. \ref{fig:position}(e)) around a single SiV(2) center. The inset shows a close up of the SiV(2) ZPL at \unit[726.0]{nm}, that red shifts by \unit[2.6]{nm} upon FIB etching. By the milling process, a new SiV(3) center (ZPL at \unit[739.9]{nm} marked by $\star$, PSB marked by $\blacktriangledown$) is created. At wavelengths \unit[750-830]{nm} M7-cavity modes are visible. (b) Cavity tuning of modes $e1$, $e3$, $o1$ and $o2$ into resonance with the SiV(3) ZPL ($\star$) and PSB ($\blacktriangle$). On resonance with the $o1$-mode, the SiV(3) ZPL intensity is enhanced by a factor of 3.8. (c) Polarization measurement of the $o1$-mode (\textcolor[rgb]{0.5,0.5,0.5}{$\bullet$}) and SiV(3) ZPL (\textbf{\tiny{$\square$}}\normalsize). (d) Simulated $E_x$-field of $o1$-mode. The emitter's position is marked by a white circle. }

\caption{(Color) Purcell enhancement of the SiV(4) ZPL: (a) Polarization polar plot of SiV(4) ZPL (\textbf{\tiny{$\square$}}\normalsize) and the M7-cavity $o2$-mode (\textcolor[rgb]{0.5,0.5,0.5}{$\bullet$}). (b) Tuning spectra of $o2$-mode: In resonance with the SiV(4) ZPL an intensity enhancement of 19 is observed. (c) On (\textcolor[rgb]{1,0,0}{$\star$}) and off (\textbf{\tiny{$\square$}}\normalsize) resonance with the cavity, the parameters $\tau_1$, $\tau_2$, $a$ are determined by fitting $g^{(2)}(\tau)$ function measured at various excitation powers $P$ with equation (\ref{eq:g2}). $P$ is normalized to the saturation power $P_{\mathrm{sat,on}} = $ \unit[0.89]{mW} and $P_{\mathrm{sat,off}} = $ \unit[0.98]{mW} on- and off-resonant with the cavity. Solid lines: theoretical power dependance of $\tau_1$, $\tau_2$, $a$. (d) $E_x$-field of the $o2$-mode. The emitter's position is marked by a white circle.}

\caption{(Color) Continuous polarization control: Polarization emission angle $\Phi$ of the SiV phonon side band (\tiny$\blacksquare$\normalsize) as a function of the detuning from the cavity modes $m_{0^{\circ}}$ and $m_{-45^{\circ}}$ with polarization angles $\Phi = 0^{\circ}$ (\textcolor[rgb]{1,0,0}{$\bullet$}) and $\Phi = -45^{\circ}$ (\textcolor[rgb]{0,0.5,0}{$\blacktriangle$}), respectively. The solid lines are a guide to the eye. On the right: Polar plots of SiV phonon side band (\tiny$\blacksquare$\normalsize) and of cavity modes (red/green solid line) for selected detunings (\footnotesize$\Box$\normalsize).}

147 — 1408.2391

\caption{Collapse due to bridging proteins. A: Switch-like transition to a collapsed rosette-like state for polymers for small numbers of bridging interactions $p$, solid lines (---) show the predictions of the theoretical estimates assuming two states (Eq.\ref{eq:Zloop_red}), while red dots (\textcolor{red}{$\cdots$}) correspond to the full estimate for $p=4$ (Eq.~\ref{eq:prob_collapseZ}), both taking into account the measured scaling $P_\mathrm{(loop)} \sim g^{-2.27}$. The star polymer contribution has been neglected (see Fig.~\ref{fig:6}). % B: Two-state dynamics of the number of active bridging interactions (contacts) for $p = 4$. The interaction energy $\epsilon_l = 6.52 k_B T$ is set close to the critical value measured in panel A. The plot shows switching between a compact and a swollen state as a function of Monte Carlo time, compatibly with a first-order phase transition showing phase coexistence between a completely collapsed state (with six contacts) and the swollen state. }

148 — 1408.4154

\caption{\textcolor{red}{Da \ref{fig:prisoner_shorttime_1} a \ref{fig:prisoner_shorttime_3}} \textbf{Prisoner's Dilemma: short time behaviour}. Plot of the solution $u_{\gamma_0,\gamma_1}(x,T)$ for a fixed short time $T$ and several values of $\gamma_0,\gamma_1$, varying the value of $m_0,m_1$. Figures (a) and (b) refer to fair mutation, so $(m_0,m_1)$ is in the extinction region $E$. %; precisely $(m_0,m_1)=(0,0.4)$ in (a), and Figures (c) and (d) refer to unfair mutation: in (c) $(m_0,m_1)$ is in the coexistence region $C_0$, while in (d) $(m_0,m_1)$ is in the fixation region $F$. Figures (e) and (f) refer to back and forth mutations, so $(m_0,m_1)$ is in the coexistence range $C_1$. %; precisely $(m_0,m_1)=(0.1,0.1)$ in (e), and$(m_0,m_1)=(0.9,0.9)$ in (f). We see that for small values of $m_0+m_1$, the graph of $u_{\gamma_0,\gamma_1}$ stays above the replicator-mutator's one and increases with $\gamma_0, \gamma_1$ (figures (a), (c), (e)). The situation is reversed for large values of $m_0+m_1$ (see figures (b), (d), (f)). }

149 — 1408.4621

\caption{ An overview of the field-of-view (FOV) inferred by SST/CRISP and SDO/AIA consisting of: (a) SDO/AIA $170$ nm, detailing the photosphere; (b) SST H$\alpha$ $656.28$ nm (line core) sampling the chromosphere; the (c) SDO/AIA $30.4$ nm filter (TR); and the lower corona detailed by (d) SDO/AIA $17.1$ nm. The white line on each image represents the slit used to construct the time-distance diagrams plotted in Fig. \ref{fft_slit}. The yellow and cyan lines outline each slit used to investigate UPW behaviour. \textcolor{red}{\textbf{The yellow slits show where UPWs were observed and cyan slits show no UPWs.}} }

\caption{ The base layer indicates the magnetic field inferred by the SDO/HMI instrument. The purple box highlights the SST/CRISP FOV which is overlaid. An extended FOV context image from the SDO/AIA $30.4$ nm filter is also included. The green lines are the visualisation of the magnetic field extrapolation. A strong correlation exists between these lines and the brighter regions in the SDO/AIA $30.4$ nm image underpinning that the extrapolation is a \textcolor{red}{\textbf{reasonable approximation over such a large height.}} }

\caption{ (Top row) Unfiltered time-distance slits for the H$\alpha$ line core (a), SDO/AIA $30.4$ nm filter (b), and $17.1$ nm filter (c) constructed for the white slit in Fig. \ref{overview}. (Middle row) Time-filtered 3-minute FFT output for H$\alpha$ (d), SDO/AIA $30.4$ nm (e), and SDO/AIA $17.1$ nm (f). (Bottom row) 5-minute FFT output for H$\alpha$ (g), SDO/AIA $30.4$ nm (h), and SDO/AIA $17.1$ nm (i). The windows used are centred on $3\pm1.5$ mhz (referred to as $5$ minutes) and $5\pm1.5$ mhz (referred to as $3$ minutes). \textcolor{red}{\textbf{The white dotted line is the pore boundary, below the line is the pore and above is the background chromosphere.}} }

150 — 1408.4802

\caption{Rayleigh number ranges over which we have found shearing ($*$) and non-shearing ({\color{blue}{\large$\circ$}}) convection to persist with $A=2$ and $Pr=1$, 3, or 10. Upper bounds on the non-shearing regimes ($Ra=1.2\e6$, $3.5\e6$, and $1.8\e7$) and the lower bounds on the shearing regimes ($Ra=2.5\e4$, $3\e5$, and $2\e6$) are only approximate. The shearing flows are \emph{bursting} when $\Pran=1$ and \emph{sustained} when $\Pran=3$ or 10 (see text).}

\caption{Time-averaged Nusselt numbers ($N$), total kinetic energies ($E$), and the fractions of kinetic energies due to vertical velocities ($E_z/E$) in sustained shearing convection ($*$) and non-shearing convection at various Rayleigh numbers for $(A,\Pran)=(2,10)$. The non-shearing solutions seem to be variously chaotic (\textcolor{blue}{\tiny{$\square$}}), periodic (\textcolor{blue}{\small{$\diamond$}}), and quasiperiodic (\textcolor{blue}{\small{$\triangleleft$}}), while the shearing solutions all seem to be chaotic. Temperature fields from the cases labelled $a$-$d$ are shown in figures \hbox{\ref{fig: video stills}(a-d)}, respectively.}

\caption{Mean vertical profiles of zonal flow (left) and temperature (right) for sustained shearing convection with $(A,\Pran)=(2,10)$ at Rayleigh numbers of $2\cdot10^6$ (\protect\dashdotrule), $2\cdot10^7$ (\protect\dashedrule), and $2\cdot10^8$ (\solidrule). Each zonal flow profile is normalized by its maximum value, $\overline{u}^t_{max}$. Reversing $\overline{u}^t(z)$ yields the profiles that arise when symmetry breaks in the opposite way. Figure \ref{fig: Pr=1 profiles} shows analogous profiles for \emph{bursting} shearing convection.}

\caption{Mean vertical profiles of zonal flow (left) and temperature (right) for bursting shearing convection with $(A,\Pran)=(2,1)$ at Rayleigh numbers of $2\e6$ (\protect\dashdotrule) and $2\e8$ (\solidrule). Each zonal flow profile is normalized by its maximum value, $\overline{u}^t_{max}$. Reversing $\overline{u}^t(z)$ yields the profiles that arise when symmetry breaks in the opposite way. Figure \ref{fig: Pr=10 profiles} shows analogous profiles for \emph{sustained} shearing convection.}

\caption{Mean Nusselt numbers, $N$, of shearing convection at three Prandtl numbers and various Rayleigh numbers in a domain with $A=2$. The shearing convection is \emph{bursting} with $\Pran=1$ (\textcolor[rgb]{0,.6,0}{\scriptsize{$\blacktriangle$}}) and \emph{sustained} with $\Pran=3$ (\textcolor[rgb]{.8,0,0}{\scriptsize{$\bullet$}}) and $\Pran=10$ ($*$). In the sustained cases, dashed lines show power-law fits of $N\sim 3.4\,Ra^{0.077}$ for $\Pran=3$ and $N\sim 1.0\,Ra^{0.19}$ for $\Pran=10$. These $N$ values are compared to those of non-shearing states for $\Pran=10$ and $\Pran=1$ in figures \ref{fig: Pr=10 int} and \ref{fig: Pr=1 N}, respectively.}

\caption{Mean Nusselt numbers, $N$, for various Rayleigh numbers with $\Pran=1$. For $A=2$, bursting shearing convection (\textcolor[rgb]{0,.6,0}{\scriptsize{$\blacktriangle$}}) and non-shearing convection (\textcolor{blue}{\scriptsize{$\blacktriangledown$}}) are represented. For $A=1$, only bursting shearing convection (\textcolor[rgb]{.8,0,0}{\tiny$\triangle$}) is represented. Nusselt numbers computed by \citet{Johnston2009} for \emph{no-slip} boundaries and $A=2$ are also shown ({\large$\circ$}).}

151 — 1408.5298

\caption{% (Color online) Covariances for Gaussian boundary conditions with parameters as in Figure~\ref{fig:Gaussian1}. The initial and final covariance matrices of $\boldsymbol{x}\equiv \left[\boldsymbol{q}\,,\boldsymbol{p}\right]$ are given by $\begin{psmallmatrix} 1 & 0 \\ 0 & 1 \end{psmallmatrix}$ and $\begin{psmallmatrix} 1.7 & 0 \\ 0 & 1 \end{psmallmatrix}$, \resp } \label{fig:Gaussian2} \end{figure} As $g \to 0$ the limit behavior of the cumulants is described by the ``slow manifold'' specified by the condition $\mathsf{V}^{\scriptscriptstyle{(1)}}=0$ and the evolution law (\ref{algebraic}). Only in a layer close to the boundaries of the control horizon cumulants get away from the slow manifold along exponentially stable and unstable directions with rates of the order $O(1/\sqrt{g})$ in order to satisfy the boundary conditions. The description of singular boundary value problems in terms of invariant manifolds is well known in the theory of dynamical systems \cite{TiKoJo94}. We refer the interested reader to \cite{PMGSc14b} for the details of the multiscale expansion \cite{PaSt08} proving the foregoing qualitative picture. \section{Overdamped limit} The relevance of the ``slow manifold'' condition $\mathsf{V}^{\scriptscriptstyle(1)}=0$ appears from the fact that it permits to recover directly at $g=0$ the ``overdamped'' limit of the Langevin--Kramers dynamics. The overdamped regime corresponds to the assumption of a wide scale separation between the control horizon $[0,\tf]$ and the Stokes time $\tau$ and between the characteristic length scale $L$ of the configuration space boundary data $ \mu_{\mathrm{\iota}}(\boldsymbol{q}/L)$, $\mu_{\mathrm{f}}(\boldsymbol{q}/L)$ and the typical length scale $\ell=\tau/\sqrt{\beta\,m}$ of the uncontrolled process. In the overdamped regime, the quantifier of the scale separation is the Stokes number $\st=\tau/\tf=\ell^{2}/L^{2}\ll\,1$. Under these hypotheses, we can look for an asymptotic solution of (\ref{FP}) and (\ref{dp}) by expanding around a Maxwell--Boltzmann momentum equilibrium distribution $\mu_{\scriptscriptstyle{MB}}(\boldsymbol{p})$ perturbed at large scales, $(\boldsymbol{\tilde{q}},\tilde{t})\equiv(\sqrt{ \st}\,\boldsymbol{q},\st t)$, by the action of a control potential of the form $U(\boldsymbol{q},t)\equiv U_{0}(\sqrt{ \st}\,\boldsymbol{q},\st t)+O\left(\sqrt{ \st}\right)$. From now on we specify by an underscript the order of the perturbative expansion. Upon setting $\partial_{\boldsymbol{\tilde{q}}}S_{0}=-\partial_{\boldsymbol{\tilde{q}}}\ln \mu_{0}$, and applying standard homogenization techniques (see e.g. \cite{PaSt08}, see also \cite{PMG14}) we get for the solution of the Fokker--Planck equation \eqref{FP}: \begin{eqnarray} \label{centeringFP} &&\rho(\boldsymbol{p},\boldsymbol{\tilde{q}},\tilde{t})= \mu_{\scriptscriptstyle{MB}}(\boldsymbol{p})\,\mu_{0}(\boldsymbol{\tilde{q}},\tilde{t})\times \nonumber\\&& \Bigl( 1 + \sqrt{ \st}\,\tfrac{\tau}{m} \,\boldsymbol{p} \cdot \partial_{\boldsymbol{\tilde{q}}} (S_{0} - \beta\,U_{0} )(\boldsymbol{\tilde{q}},\tilde{t}) +O(\st) \Bigr) \end{eqnarray} For the solution of the dynamic programming \eqref{dp} we obtain \begin{eqnarray} \label{centeringV} &&V(\boldsymbol{p},\boldsymbol{\tilde{q}},t,\tilde{t}) = \frac{(\tf-t)\,d}{\beta\,\tau} + \frac{\|\boldsymbol{p}\|^{2}}{2\,m} + V_{0}(\boldsymbol{\tilde{q}},\tilde{t}) + \nonumber\\&& \sqrt{\st} \Bigl( V_{1}(\boldsymbol{\tilde{q}},\tilde{t}) + \tfrac{\tau}{m} \,\boldsymbol{p} \cdot \partial_{\boldsymbol{\tilde{q}}}\,(V_{0}-U_{0})(\boldsymbol{\tilde{q}},\tilde{t}) \Bigr) +O(\st) \end{eqnarray} The functions $S_{0}$ in \eqref{centeringFP} and $V_{0}$ in \eqref{centeringV}, respectively, obey the local equilibrium potential equation \begin{equation*} %\label{cellFP} \partial_{\tilde{t}}S_{0}-\tfrac{\tau}{m} \Bigl( (\partial_{\boldsymbol{\tilde{q}}}S_{0}) \cdot\partial_{\boldsymbol{\tilde{q}}} - \partial_{\boldsymbol{\tilde{q}}}^{2} \Bigr) ( U_{0} - \tfrac{1}{\beta}S_{0} ) = 0 \end{equation*} and the dynamic programming equation \begin{equation*} %\label{cellV} \partial_{\tilde{t}}V_{0}-\tfrac{\tau}{m} \Bigl( (\partial_{\boldsymbol{\tilde{q}}}U_{0}) \cdot \partial_{\boldsymbol{\tilde{q}}} - \tfrac{1}{\beta} \partial_{\boldsymbol{\tilde{q}}}^{2} \Bigr) (V_{0}-U_{0})=0 \end{equation*} These equations specifying two of the three optimal control equations governing the minimal heat release by a Langevin--Smoluchowski dynamics between $\mu_{\iota}$ and $\mu_{\mathrm{f}}$ \cite{AuMeMG11}. In order to recover the third condition, we use \eqref{centeringFP} and \eqref{centeringV} to evaluate \begin{equation*} \mathsf{V}^{\scriptscriptstyle{(1)}}(\boldsymbol{\tilde{q}},\tilde{t}) = -\sqrt{\st} \,\tfrac{\tau}{m} \,\partial_{\boldsymbol{\tilde{q}}} \Bigl( 2 \, U_{0} - \frac{S_{0}}{\beta} - V_{0} \Bigr) (\boldsymbol{\tilde{q}},\tilde{t})+O(\varepsilon) \nonumber \end{equation*} Then the condition $\mathsf{V}^{\scriptscriptstyle{(1)}}=0$ yields exactly the relation between $U_{0}$, $V_{0}$, $S_{0}$ that allows us to recover the very same Monge--Amp\`ere--Kantorovich equations of\cite{AuMeMG11}: \begin{gather*} \partial_{\tilde{t}}\tilde{U} -\tfrac{\tau}{2\,m}\partial_{\boldsymbol{\tilde{q}}}\tilde{U}\cdot\partial_{\boldsymbol{\tilde{q}}}\,\tilde{U} %\end{eqnarray} %\begin{eqnarray} \\ \partial_{\tilde{t}}S_{0} - \tfrac{\tau}{m} (\partial_{\boldsymbol{\tilde{q}}}S_{0})\cdot\partial_{\boldsymbol{\tilde{q}}} \tilde{U} + \tfrac{\tau}{m} \partial_{\boldsymbol{\tilde{q}}}^{2} \tilde{U} = 0 \end{gather*} with $\tilde{U}\equiv U_{0}-S_{0}/\beta$. \section{Conclusion} We showed how Pontryagin's principle can be used to derive refined bounds for the Second Law of thermodynamics in the case of nanomechanical systems. We also established a relation between optimal control and kinetic theory, which renders available ideas and tools of dilute gas \cite{CeIlPu94,Lev96} and optimal transport theory \cite{Villani} to the construction of optimal protocols implementing at the nano-scale information processing operations such as the erasure of a bit \cite{DiLu09}. The authors acknowledge the Finnish Academy Center of Excellence ``\emph{Analysis and Dynamics}'' for support. The authors are grateful to Carlos Mej\'{i}a-Monasterio and Antti Kupiainen for helpful comments. \addcontentsline{toc}{section}{Bibliography} \bibliography{/home/paolo/RESEARCH/BIBTEX/jabref}{} %\bibliographystyle{aipauth4-1} %base bibstyle abbrv.bst with eprint field \bibliographystyle{aipnum4-1} \end{document} }

152 — 1408.5850

\caption{\label{3danderson}[Color online] Entanglement Entropy divided by square of linear system size of three dimensional Anderson model in log-linear scale for different value of disorder strength $w$. $E_F=0$. Mean free path corresponding to selected $w$'s is indicated as a vertical line. Data points corresponding to $w_c=16$ are indicated by \textcolor{blue}{$\blacksquare$}. Horizontal axis is the linear size of the system, $L$. The total system has $L \times L \times L$ sites. Number of samples is $100$ for small sizes and $10$ for large sizes.}

153 — 1408.6086

\caption{Optimization of a fast readout pulse. \textcolor{blue}{(a)} initial pulse sequence with fidelity $\Phi_\text{ch,i}'=83.8\%$ and contrast $\xi_\text{i}=19.8\%$. \textcolor{blue}{(b)} Optimized pulse shape. \textcolor{blue}{(c)} Initial time evolution of populations. Again, the unoptimized pulse fails to let $\ket{1}$ tunnel into $\ket{m}$. \textcolor{blue}{(d)} Time evolution of populations after pulse optimization resulting in a high contrast of $\xi_\text{f}=98.2\%$ and final fidelity $\Phi_\text{ch,f}'=99.2\%$. \label{Fig:PII_Ch6_TFastPulses}}

154 — 1408.6370

\caption{Loss function $-\Im\{\varepsilon^{-1}(\textbf{q},\omega)\}$ versus energy $\hbar\omega$ in eV from TDDFT-RPA calculations for $\|\textbf{q}\|= q_y\approx 0.1$~\AA$^{-1}$ and $\gamma \approx 0.5$~eV, from standard DFT with $L_z \approx 10$~\AA\(-----), $L_z \approx 40$~\AA\({\color{red}{\textbf{------}}}), augmented with zero-padding so $L_z \approx 10 + 30 \approx 40$~\AA\(--~--~--), and including a radial cutoff so $L_z \approx 10 + 10 \approx 20$~\AA, $R \approx L/2 \approx 10$~\AA\({\color{blue}{\textbf{--~$\cdot$~--~$\cdot$}}}). Experimental data from Ref.~\cite{exp1} ($\color{Tan}{\bullet}$) is provided for comparison.}

\caption{Plasmon energies in eV versus unit cell parameter $L_z$ in \AA\for the (a)$\sigma+\pi$ plasmon $\omega_{\sigma+\pi}$ and (b) $\pi$ plasmon $\omega_{\pi}$, obtained from TDDFT-RPA calculations of $\max_{\omega}-\Im\{\varepsilon^{-1}(\textbf{q},\omega)\}$ for $\|\textbf{q}\| = q_y \approx 0.1$~\AA$^{-1}$ (insets), from standard DFT ({$\medbullet$},\textbf{------}), augmented by zero-padding ({\color{red}{$\blacksquare$},\textbf{--~--~--}}), and including a radial cutoff of $R \approx L_z/2$ ({\color{blue}{$\Diamondblack$}}). Grey regions denote experimental range of $\omega_{\sigma+\pi}$ and $\omega_{\pi}$ \cite{exp1}. }

155 — 1409.0199

\caption{(Color Online)(a) Chronophotography of the transient motion of the walker being trapped by the magnetic field. The time increment is fixed at 0.05 s which corresponds to one point every two bounces. After a few oscillations, the walker trajectory converges to a circular orbit. (b) Transient regimes obtained with Fort's model of the walker dynamics (solid black line) and numerical solving of the Rayleigh equation with $\Gamma=25.5$ [solid red line (online), light gray (printed)]. The quantitative agreement between experiments and numerical simulations shows that the walker propulsion can be described through a friction term depending on the velocity. (c) Orbit radius $R/\lambda_F$ as a function of the dimensionless frequency $\lambda_F \omega / V$, where $\omega$ is the characteristic frequency of the drop in the harmonic well. Experiments for two drops of velocity $V$=10mm/s and 8 mm/s (\color{red}{$\circ$}).\color{black} The experimental data (red dots online) are compared to the simple scaling $R/\lambda_F=V/\omega\lambda_F$ (solid black line downer). The agreement is good, without use of any fit parameter. The blue (upper) line (blue online, gray printed) indicates the result of Fort's numerical model. \label{fig2}}

156 — 1409.0602

\caption{Basic idea of the proposed transductive alignment. The goal is to automatically label source-type landmarks (\textcolor{magenta}{magenta}) under the guidance of common landmarks (\textcolor{blue}{blue}).}

157 — 1409.0818

\caption{(Color online) Results of the model calculation showing the transition between positive and negative correlations between the conductance variations on a resonance crossing. If not stated otherwise in the subfigures, the tunnel couplings are set to \textcolor{blue}{$\Gamma_{S1}=0.01$, $\Gamma_{N1}=0.1$}, \textcolor{red}{$\Gamma_{S2}=0.005$, $\Gamma_{N2}=0.05$} and $\Gamma_{12}=0.001$. (a) and (b) Transition induced by tuning $\Gamma_{N1}$, (c) and (d) transition induced by tuning $\Gamma_{S2}$, and (e) and (f) transition induced by tuning $\Gamma_{12}$, for fixed to $\Gamma_{S2}=0.017$.}

158 — 1409.1238

\caption{IRAC \mcolor{3.6}{4.5} vs. \mcolor{4.5}{5.8} color-color plot indicating the region occupied by PDRs (PAH grains and UV fluorescent H$_2$). Region 1 of color-color space is occupied by gas clouds of thickness $A_{V} \sim$ 0--2 mag. Region 2 of color-color space is occupied by either a gas cloud of thickness $A_{V} \sim$ 2--4 mag in front of the incident UV source or a gas cloud ($A_{V} > 2$) behind the incident UV source. Region 3 indicate a cloud thickness $A_{V} \sim$ 4--5 mag in front of the incident FUV source. The grey region at \mcolor{4.5}{5.8} $>$ 3.6 indicates region of color space with significant degeneracy. }

\caption{{\it Spitzer} IRAC \mcolor{3.6}{4.5} vs. \mcolor{4.5}{5.8} color-color plot showing the location of shocked H$_2$ emission in relation to UV excited PDR emission.}

159 — 1409.1952

\caption{(Color online) {\bf Average infidelity as a function of number of measurements for a single-qubit state estimation protocol.} Four strategies are compared: (1) the proposed protocol ({\color{blue}$\circ$}), (2) a restricted-adaptive strategy ({\color{red}$\times$}), (3) a restricted-nonadaptive strategy with a fixed set of bases ({\color{ForestGreen}$\triangle$}) and (4) a random-basis strategy in which the measurement bases are chosen at random from the Haar-measure at each iteration ({\color{Orange}$\blacksquare$}). The bars indicate standard errors of the mean. The solid line is the theoretical bound achievable by an optimal collective measurement scheme on the entire ensemble of qubits. The inset shows the difference between the mean infidelity and the theoretical bound~\cite{massar95}, as a function of the number of measurements, for the various methods. The figure clearly illustrates the advantage that the proposed method has over others to quickly approach the theoretical bound.}

\caption{(Color online) {\bf Average infidelity as a function of the number of measurements for a two-qubit state estimation protocol.} Two strategies are compared: (1) a restricted-adaptive strategy ({\color{red}$\times$}), and (2) a restricted-nonadaptive strategy ({\color{ForestGreen}$\triangle$}). The error bars are smaller than the symbols. Even on a restricted set of measurement bases, we find an appreciable advantage in using the adaptive strategy over the nonadaptive one.}

160 — 1409.2845

\caption{Sketch of the studied heterostructure showing the 530 nm thick Finemet\textregistered{} film deposited onto a 125 $\text{\textmu}$m thick Kapton\textregistered{} substrate and glued onto the piezoelectric actuator. The arrows qualitatively represent the in-plane strains of the piezoelectric actuator.}

\caption{Brillouin light scattering spectrum of the amorphous Finemet\textregistered{} film on Kapton\textregistered{} substrate. The transferred wave-vector is $q=2.14\times10^{5}$ cm$^{-1}$. $R$ denotes Rayleigh surface wave while $S_{i}$ correspond to the Sesawa guided waves. The red dashed line corresponds to the theoretical calculated spectrum while the continuous blue line corresponds to the experimental data.}

161 — 1409.3408

\caption{(a)-(c) images of the predictive distribution of one missing locations, obtained from 200 hundred locations generated uniformly over the unit square and simulated data from a Guassian process with $\phi=0.1$ (a) and $\phi=0.5$ (c); (b)-(d) images of $95\%$ high density regions for the predictive distributions in (a) and (c), repsectively, showing the true location (o), the mean (\protect\marksymbol{triangle}{black}), the mode ($+$) and the componentwise median ($\times$).\label{fig:example_simulation}}

\caption{(a)-(b) images of the predictive distribution of one missing locations, obtained from 200 hundred locations generated from an inhibitory point process with minimum distance between locations $\delta = 0.04$ (a) and $\delta = 0.06$ (b); the true location (o), the mean (\protect\marksymbol{triangle}{black}), the mode ($+$) and the component-wise median ($\times$) are also shown.\label{fig:example_inhibition}}

162 — 1409.3987

\caption{Left: 21~cm \textsc{Hi} absorption map of UCG~7408 corresponding to the channel at velocity, $v=444.5$~\kms. The colors show the observed flux densities such that blue represents pixels with negative flux density i.e. {\color{blue}absorption}, yellow/red represents pixels with positive flux density i.e. {\color{red} emission}, and green indicate pixels within the observed {\color{green} noise ($\pm 1\sigma$)} in the data. The background source is shown in black contours at flux density levels of 5, 10, 20, 30, 40, and 50 mJy/beam. It is worth noting that we are sensitive to absorption only against the background radio source. However, we are sensitive to emission in the entire region. The synthesized beam size ($\equiv$230~pc$\times$173~pc in physical units) is shown in the lower left-hand corner. The red contours show the VLA D-configuration \textsc{Hi} emission map with contour levels indicating 14.4, 21.7, and 28.9 $\rm \times 10^{19}~cm^{-2}$. Spectra extracted from the regions marked in white are shown in the right panel. Right: VLA B-configuration \textsc{Hi} spectrum extracted from the region marked in the white oval, the size of the beam, is shown in the top panel in black. The absorption has a peak depth of 10.41~mJy and the flux density of the background source at the same region is 13.98~mJy. The absorption feature is unresolved and the limiting FWHM of was measured to be 1.1~\kms. This corresponds to a kinetic temperature, $\rm T_k\le$26~K. The VLA-D configuration \textsc{Hi} spectrum extracted from the white dashed rectangular region (similar in size to the D-configuration beam) is shown in the lower panel in blue. The feature has a FWHM of 4.75~\kms and a centroid at 442.2~\kms. The absorption spectrum was obtained by subtracting out the Gaussian emission profile from the raw spectrum as discussed in detail by B11. }

163 — 1409.4705

\caption{\label{fig:MultiClk2-4lens-UR} \textbf{Example of a practical paraxial cloak.} \textbf{(a)-(c)} A hand is cloaked for varying directions, while the background image is transmitted properly (See \textcolor{blue}{Media~1,2} for videos). \textbf{(d)} On-axis view of the ray optics cloaking device. \textbf{(e)} Setup using practical, easy to obtain optics, for demonstrating paraxial cloaking principles. (Photos by J. Adam Fenster, videos by Matthew Mann / University of Rochester)}

\caption{\label{fig:MultiClk2-4lens-exp} \textbf{Experimental demonstration of a `perfect' paraxial cloak with four lenses.} Camera was focused on the wall. The grids on the wall can be seen clearly, and match the background for all colors and viewing angles. The middle of the ruler is cloaked inside the lens system for all angles shown. Images at various camera-viewing angles: \textbf{(a)} $-0.65^\circ$, \textbf{(b)} on-axis ($0^\circ$), \textbf{(c)} $0.47^\circ$, \textbf{(d)} $0.95^\circ$. \textbf{(e)} Side profile of experimental setup. See \textcolor{blue}{Media~3,4} for videos. (Videos by Matthew Mann / University of Rochester) }

164 — 1409.4995

\caption{\textbf{Tagging results produced by our image annotation system}. Tags uniquely selected by the adaptive tag selection method are shown in an {\color{blue}\textit{italic}} font.}

165 — 1409.5597

\caption{Evolution of the global binary fraction and, for the \nbody\model, the fraction in the core\red{(dots)}. For this plot, data was lost around 6 Gyr. Core data is not readily available for the Monte Carlo model. }

\caption{Line-of-site velocity dispersion in the \nbody\model, compared with the observational data of\citet{Pe1995}.}

\caption{V luminosity functions in annuli with median radii of 0.938 (top) and 5.028 arcmin (bottom). The \nbody\model is compared with the observational data of\citet{Ri2004} \red{and the Monte Carlo results of \citet{2008MNRAS.389.1858H}}.}

166 — 1409.6036

\caption{Constraint on CP violating topological $\theta$ parameter for discrete integer values of {\it Pontryagin index} $Q_P$ using {\tt Planck + WMAP9} best fit cosmological parameters. Here \textcolor{red}{Red} and \textcolor{black}{\bf black} colored points correspond to the upper and lower bound of the $\theta$ parameter for a given value of $Q_P$. All the parallel \textcolor{blue}{blue} colored lines are drawn for different integer values of $Q_P$ which connects both the \textcolor{red}{Red} and \textcolor{black}{\bf black} colored points. This plot suggests that as the value of $Q_P$ increases then the interval between the upper and lower bound of the $\theta$ parameter decrease and it will converge to very small value for large $Q_P$. Also the numerical value corresponding to the upper bound and lower bound of the $\theta$ parameter decreases once we increase the the value of $Q_P$. }

167 — 1409.6352

\caption{\small % A portion of a bounded packing, and $C_0$ is one % of the inner circles. The part of ${\mathcal P}_0$ we illustrated is shaded % in colors, and $C_0$ is in \textcolor{blue}{blue}. $C_\epsilon$ \is in dashed % \textcolor{blue}{blue}. A portion of a bounded packing. Here $C_0$ is an arc of a \textcolor[rgb]{0.7,0.6,0}{yellow} disk. The circles of ${\mathcal P}_0$ are in various shades of \textcolor{red}{red}, $C_0$ is in \textcolor{blue}{blue} and $C_\epsilon$ is in dashed \textcolor{blue}{blue}. }

\caption{\small A portion of a bounded packing, where $C_0$ is the outer circle. The circles of ${\mathcal P}_0$ are in various shades of \textcolor{red}{red}, $C_0$ is in \textcolor{blue}{blue} and $C_\epsilon$ is in dashed \textcolor{blue}{blue}.}

\caption{\small A portion of a bounded packing, where $C_0$ is an inner circle. The circles of ${\mathcal P}_0$ are in various shades of \textcolor{red}{red}, $C_0$ is in \textcolor{blue}{blue} and $C_\epsilon$ is in dashed \textcolor{blue}{blue}. }

168 — 1409.6755

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Left panel:} shows the distribution of GRB rate in the density-temperature ($n-T$) parameter space. Note that the density and temperature are defined on the local gas cell of scale of a few pc that a GRB sits in and it will be made clear later that the appearance of GRB afterglows is in most cases more dependent on the properties of gas along the line of sight. We have further divided the GRBs into two groups with respect to intervening neutral hydrogen column density: $N_{HI}>10^{19}$cm$^{-2}$ (red) and $N_{HI}<10^{19}$cm$^{-2}$ (blue), details of which will be given in subsequent figures. The contour levels specified indicate the fraction of GRBs enclosed. {\color{burntorange}\bf Right panel:} shows the distribution of GRB rate in the density-metallicity ($n-Z$) parameter space. }

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Left panel:} shows the PDFs of total hydrogen column density for the three sub-populations of GRBs: GRBs in $(n_H,T)=(10^{-2.5}{\rm cm}^{-3},10^{7.5}{\rm K})$ with $N_{\rm HI}\le 10^{19}$cm$^{-2}$ (red dashed, LnLN group), $(n_H,T)=(10^{-2.5}{\rm cm}^{-3},10^{7.5}{\rm K})$ with $N_{\rm HI}\ge 10^{19}$cm$^{-2}$ (green dotted, HnLN group), and $(n_H,T)=(10^{4.0}{\rm cm}^{-3},10^{3.75}{\rm K})$ with $N_{\rm HI}\ge 10^{19}$cm$^{-2}$ (blue solid, HnHN group). {\color{burntorange}\bf Right panel:} shows the cumulative PDFs of the ratio of $N_{HI}/N_{H}$. }

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Left panel:} shows the PDFs of metallicity weighted total hydrogen column density, $(N_{H}/10^{21}{\rm cm}^{-2})(Z/\zsun$, excluding gas with temperature greater than $10^6$K, for the three sub-populations of GRBs: GRBs in $(n_H,T)=(10^{-2.5}{\rm cm}^{-3},10^{7.5}{\rm K})$ with $N_{\rm HI}\le 10^{19}$cm$^{-2}$ (red dashed), $(n_H,T)=(10^{-2.5}{\rm cm}^{-3},10^{7.5}{\rm K})$ with $N_{\rm HI}\ge 10^{19}$cm$^{-2}$ (green dotted), and $(n_H,T)=(10^{4.0}{\rm cm}^{-3},10^{3.75}{\rm K})$ with $N_{\rm HI}\ge 10^{19}$cm$^{-2}$ (blue solid). The exclusion of $\ge 10^6$K gas is to intended for the situation that dust is efficiently destroyed in hot gas. According to \citet[][]{2003Draine}, $A_V\approx (N_{H}/10^{21}{\rm cm}^{-2})(Z/\zsun$. {\color{burntorange}\bf Right panel:} shows the PDFs of gas temperature weighted by $N_{H}Z$, excluding gas with temperature greater than $10^6$K. }

169 — 1409.7592

\caption{Sketch of pseudo-2D random dimers: (a) {\sc ecoradi} and (b) {\sc dicoradi}. The individual realizations of $\epsilon_n$ and $\epsilon_m$ are shown on both axis. \tikzcircle[fill=white]{3.4pt} $\epsilon_{n,m}=0$, \tikzcircle[fill=lightgray]{3.4pt} $\epsilon_{n,m}=\Delta$, \tikzcircle[fill=gray]{3.4pt} $\epsilon_{n,m}=2\Delta$.}

170 — 1409.7598

\caption{Observed cognitive measures (IST15 in {$\Diamondblack$} and WST in {\textcolor{blue}{$\bullet$}} ) and cause-specific individual predictive cumulative incidences (of dementia in black and death in grey) up to an horizon of 5 years for a ApoE- man who graduated from primary school and entered the cohort at 68 years old. Predicted cumulative incidences are computed using a Monte-Carlo approximation with 2000 draws. Are given the median and the 2.5\% and 97.5\% percentiles.}

171 — 1409.7985

\caption{Plate diagrams for the IP models (left) and random utility model (right). $\mathcal{J}$, $\mathcal{C}$ and $\mathcal{A}_i$ are the sets of justices, cases, and amicus briefs (for case $i$), respectively. $\bm\psi_j$ is the IP for justice $j$; $\bm\kappa_i$ is the set of case parameters $a_i, b_i, c^\petitioner_i$ and $c^\respondent_i$ for case $i$; and $\bm\alpha, \bm\sigma, \lambda$, and $\rho$ are hyperparameters. The mixture proportion nodes (dashed) are fixed in our estimation procedure. On the left, black nodes comprise the basic IP model, \textcolor{plateblue}{blue} nodes are found in both the issues IP and amici IP models, while \textcolor{platered}{magenta} nodes are found only in the amici IP model.\label{fig:plate-diagram}}

172 — 1409.8278

\caption{ {\bf Small-worldness are different in the connectomes.} The small-world measures, characteristic path length $L$, clustering coefficient $C$, small-world index $\sigma_{\text{sw}}$, and total wiring length $W$, of human (\textcolor{red}{$\bullet$}) and {\it C. elegans} (\textcolor{blue}{$\blacksquare$}) connectomes with respect to modularity, $Q$, which is varied by link swapping. Note that $W$ is normalised with respect to the values of the original neural networks. Unobservable error bars lie within the symbols. The vertical dashed lines denote the values of the original networks. The original networks show more global segregation (higher $L$ suggests lower global efficiency) and more local integration (higher $C$ suggests higher local efficiency) at the same time.}

\caption{ {\bf Compression ratio as a function of modularity.} The compression ratio is defined as the size of the compressed network divided by the size of the original network in bytes when the networks are represented by the adjacency matrices. It is shown for the original (vertical dashed lines) and rewired networks of human (\textcolor{red}{$\bullet$}, left axis) and {\it C. elegans} (\textcolor{blue}{$\blacksquare$}, right axis). }

173 — 1409.8515

\caption{a) Example of inter-vortex distances $\delta_{Si}$ between consecutive vortices (\textcolor{red}{$\circ$}) along the downstream shear-layer mean trajectory. b) Mean inter-vortex distance $\Delta_{S}$ along the downstream shear-layer mean abscissa $S$.}

\caption{a) Pairing between the instantaneous downstream shear-layer vortices in the symmetry plane (\textcolor{blue}{$\square$}) and the closest search positions (\textcolor{green}{$\circ$}) associated with normalized phase T=0.35. b) For the same phase, reference search positions, upstream and downstream shear-layer vortices.}

174 — 1410.0383

\caption{ \label{fig:fs_pdfs} Curvature and torsion PDFs from the Gaussian random wave model and numerical simulations of random waves. (a) PDFs for curvature and torsion from simulations, analytic results and Monte Carlo integration; (b) joint PDF $P(\kappa,\tau)$ drawn from Monte Carlo integration. In (a), $P(\kappa)$ is determined by (\ref{eq:curvature_pdf}) (\textcolor{paperdarkblue}{\rule[0.4ex]{3mm}{1mm}}), and from the random wave simulations (\textcolor{paperlightblue}{\rule[0.4ex]{1.25mm}{.5mm}\rule[0.4ex]{0.3mm}{0pt}\rule[0.4ex]{1.25mm}{.5mm}}). The torsion PDF is found by Monte Carlo integration (\textcolor{paperlightgreen}{\rule[0.4ex]{0.7mm}{1mm}\rule[0.5ex]{0.3mm}{0pt}\rule[0.4ex]{0.7mm}{1mm}\rule[0.4ex]{0.3mm}{0pt}\rule[0.4ex]{0.7mm}{1mm}}) (described in Section \ref{sec:torsion}), and in the simulations (\textcolor{paperpurple}{\rule[0.4ex]{1.9mm}{0.5mm}\rule[0.4ex]{0.3mm}{0pt}\rule[0.4ex]{0.6mm}{0.5mm}}). Inset: Log-log plot indicating how $\tau$ scales for larger values, as discussed in the text of Section \ref{sec:torsion}. All Monte Carlo results come from $6\times10^{10}$ points. }

\caption{ Variation of direction cosine, curvature and torsion with respect to arclength $s$ along random vortex lines. (a) Correlation functions: direction cosine (\ref{eq:persistence}) (\textcolor{paperblue}{\rule[0.4ex]{1.9mm}{0.5mm}\rule[0.4ex]{0.3mm}{0pt}\rule[0.4ex]{0.6mm}{0.5mm}}), curvature correlation$-1$ (\ref{eq:curvaturecorrelation}) (\textcolor{paperdarkgreen}{\rule[0.4ex]{1.25mm}{.5mm}\rule[0.4ex]{0.3mm}{0pt}\rule[0.4ex]{1.25mm}{.5mm}}), signed torsion correlation (\ref{eq:torsioncorrelation}) (\textcolor{paperred}{\rule[0.4ex]{3mm}{.5mm}}) and unsigned torsion correlation$-1$ (\textcolor{paperdarkpurple}{\rule[0.4ex]{0.5mm}{.5mm}\rule[0.5ex]{0.3mm}{0pt}\rule[0.4ex]{0.5mm}{.5mm}\rule[0.5ex]{0.3mm}{0pt}\rule[0.4ex]{0.5mm}{.5mm}\rule[0.5ex]{0.3mm}{0pt}\rule[0.4ex]{0.5mm}{.5mm}\rule[0.5ex]{0.3mm}{0pt}}). (b) shows a typical short vortex segment taken from our numerical experiments, coloured with a white dot at every $0.5\lambda$ arclength distance. (c) shows the varying curvature and torsion along the segment of (b). The scale of (c) is the signed logarithm $\log_{10}(\kappa$~or~$|\tau|)\times\rm{sign}(\kappa$~or~$\tau)$, except in the shaded area which tracks the raw $\kappa$ or $\tau$ between $\pm1/\lambda$. \label{fig:fs_correlations} }

175 — 1410.0389

\caption{ AwA dataset (attributes as privileged information). The numbers are mean and standard error of accuracy over $20$ runs. % for $45$ pair of classes. % when $N=100$ training samples per class. The best result is highlighted in \textbf{boldface}, which in total is $\bf{9}$ for {\tt SVM}, $\bf{27}$ for {\tt Margin Transfer}, and $\bf{9}$ for {\tt SVM+}. Highlighted \highlight{blue} indicates significant improvement of the methods that utilize privileged information (Margin Transfer and/or SVM+) over the methods that do not (SVM). We used a paired Wilcoxon test with $95\%$ confidence level as a reference. Additionally, we also provide the SVM performance on $\Xcal^{*}$ (last column). \label{tab:AwA}}

\caption{ImageNet dataset, group of snakes (bounding box annotation as privileged information). The numbers are mean and standard error of accuracy over $20$ runs. The best result is highlighted in \textbf{boldface}. Highlighted \highlight{blue} indicates significant improvement of the methods that utilize privileged information (Margin Transfer and/or SVM+) over the methods that do not (SVM). We used a paired Wilcoxon test with $95\%$ confidence level as a reference. Additionally, we also provide the SVM performance on $\Xcal^{*}$ (last column). %The bar plots show advantage of the LUPI methods over non-LUPI (Margin Transfer versus SVM, SVM+ versus SVM). %The length of the $17$ bars corresponds to relative improvement of the accuracy over $17$ snake classes. \label{tab:ImNet_snake} }

\caption{ImageNet dataset, group of sport balls (bounding box annotation as privileged information). The numbers are mean and standard error of accuracy over $20$ runs. The best result is highlighted in \textbf{boldface}. Highlighted \highlight{blue} indicates significant improvement of the methods that utilize privileged information (Margin Transfer and/or SVM+) over the methods that do not (SVM). We used a paired Wilcoxon test with $95\%$ confidence level as a reference. Additionally, we also provide the SVM performance on $\Xcal^{*}$ (last column). \label{tab:ImNet_ball} }

176 — 1410.0470

\caption{\textup{(a)} Attempt to depict the bicausal model; \textup{(b)} a schematic showing the deterministic system (\protect\ref{eqcon})--(\protect\ref{eqmer}); the edge \protect\tikz\protect\path(0ex,0ex) edge[->] node[above=0pt, black] {$\scriptscriptstyle I$} (3ex,0ex); denotes that $P$ is the integral of $\Delta P$; see Iwasaki and Simon (\citeyear{iwasaki1994}).}

177 — 1410.1493

\caption{ Example landscapes of protein adaptation with direct selection for binding only (case 1), zoomed into the region of energy space accessible to evolutionary paths in our model. (a)~Stable protein with $\EOf = -10$ kcal/mol, (b)~marginally-stable protein with $\EOf = 3$ kcal/mol, and (c)~intrinsically-unstable protein with $\EOf = 8$ kcal/mol. In all panels $\fub = 0$, $\fuf = 1$, and $\EObone = \EObtwo = 5$ kcal/mol. The coarse-grained sequence parameters are $L = 6$ and $k=5$, with \red{effective energy matrices $\eps_\rmf$, $\eps_{\rmb_1}$, and $\eps_{\rmb_2}$ sampled from distributions that were rescaled using $L_\phys = 12$.} The black star indicates the initial state for adaptation (global maximum on $\F_1$); red triangles indicate local fitness maxima on $\F_2$, shaded according to their commitment probabilities (probability of reaching that final state starting from the initial state); black circles indicate intermediate states along paths, sized proportional to their path density (total probability of paths passing through them); small gray circles are genotypes inaccessible to adaptation. The black contours indicate constant fitness $\F_2$ (the fitness difference between adjacent contours is non-uniform so that they are equidistant in energy space), while example paths are shown in blue and green. }

\caption{ Scaling of landscape properties for three regimes of protein adaptation with direct selection for binding only (case 1). (a)~Minimum path length $\lmin$, equal to the Hamming distance between the initial and final states, versus $L(1-1/k)$; (b)~average connectivity $\gamma$ versus $L(k-1)/4$; (c)~average number $\nseq$ of accessible sequences versus $((k+1)/2)^L$; and (d)~average number $m$ of local fitness maxima versus $\nseq/(\gamma + 1)$. In all panels red circles are for stable proteins, blue squares are for marginally-stable proteins, and green triangles are for intrinsically-unstable proteins, with all energy and fitness parameters the same as in \fref{fig:case1_landscapes}. Each point represents an average over $10^4$ realizations of the folding and binding energy matrices; we exclude trivial realizations where the initial state is already a local maximum on $\F_2$. We scan over all $L > 1$ and $k > 2$ such that $k^L < 4 \times 10^4$, \red{with energy matrices rescaled using $L_\phys = 12$.} % Note that $L_\phys/L$ may not be an integer. Slope 1 lines from the origin are shown in gray to guide the eye. }

\caption{ Example landscapes of protein adaptation with selection for folding stability (cases 2 and 3). (a)~Direct selection for folding only ($\fub = 1$, $\fuf = 0$) where both the old and new targets have potentially strong binding ($\EObone = \EObtwo = 3$ kcal/mol); (b)~same selection as (a) but where the old target has weak binding ($\EObone = 10$ kcal/mol) and the new target binds strongly ($\EObtwo = 3$ kcal/mol); (c)~same selection as (a) but where the old target has strong binding ($\EObone = 3$ kcal/mol) and the new target binds weakly ($\EObtwo = 10$ kcal/mol); and (d)~direct selection for both binding and folding ($\fub = 0.9$, $\fuf = 0$) with marginal folding stability and binding strength ($\EOf = \EObone = \EObtwo = 5$ kcal/mol). All symbols are the same as in \fref{fig:case1_landscapes}. The coarse-grained sequence parameters are $L = 6$ and $k=5$, with \red{energy matrices rescaled using $L_\phys = 12$.} }

178 — 1410.1600

\caption{All Pisot numbers $\al \in \left(\tau^{4/3}, 2 \right)$ of degree $d=8$.} \begin{tabular}{l l l} \toprule no. & $\al$ & Minimal polynomial\\ \midrule 1. & $1.94284\dots$ & $x^8 - 2x^7 + x^4 - x^3 - x + 1$\\ 2. & $1.96113\dots$ & $x^8 - x^7 - x^6 - x^5 - x^4 - x^3 + 1$\\ 3. & $1.92172\dots$ & $x^8 - x^7 - x^6 - x^5 - x^4 + 1$\\ 4. & $1.94653\dots$ & $x^8 - 2x^7 + x^5 - 2x^4 + x^3 - x + 1$\\ 5. & $1.99577\dots$ & $x^8 - x^7 - x^6 - x^5 - 2x^4 + 1$\\ 6. & $1.92600\dots$ & $x^8 - x^7 - 2x^6 + x^4 - x^2 + 1$\\ 7. & $1.99203\dots$ & $x^8 - 2x^7 + x - 1$\\ 8. & $1.97061\dots$ & $x^8 - x^7 - x^6 - x^5 - x^4 - x^3 - 1$\\ 9. & $1.91743\dots$ & $x^8 - 3x^7 + 3x^6 - 2x^5 + 2x^3 - 3x^2 + 2x - 1$\\ 10. & $1.90988\dots$ & $x^8 - x^7 - x^6 - x^5 - x^4 + x + 1$\\ 11. & $1.91580\dots$ & $x^8 - 2x^7 + x^5 - x^4 - x^3 + x^2 - 1$\\ 12. & $1.96225\dots$ & $x^8 - x^7 - 2x^6 + x^4 - x^3 - x^2 + x + 1$\\ 13. & $1.97526\dots$ & $x^8 - 2x^7 + x^2 - 1$\\ 14. & $1.99402\dots$ & $x^8 - x^7 - 2x^6 - x^5 + x^4 + 2x^3 + x^2 - x - 1$\\ 15. & $1.93167\dots$ & $x^8 - 2x^7 - x^6 + 3x^5 - x^4 - 2x^3 + 2x^2 - 1$\\ 16. & $1.91451\dots$ & $x^8 - x^7 - 2x^6 - x^5 + 2x^4 + 2x^3 - x - 1$\\ 17. & $1.93895\dots$ & $x^8 - 2x^7 + x^3 - 1$\\ 18. & $1.95731\dots$ & $x^8 - 2x^6 - 3x^5 - 2x^4 + 2x^2 + 2x + 1$\\ 19. & $1.98707\dots$ & $x^8 - 3x^7 + 2x^6 + x^5 - 2x^4 + x - 1$\\ 20. & $1.99603\dots$ & $x^8 - x^7 - x^6 - x^5 - x^4 - x^3 - x^2 - x - 1$\\ \bottomrule \end{tabular}\label{pisot_d8} \end{table} \smallskip \noindent \emph{Remark 1.} Combinatorial arguments may be used to prove that the equation $\al_1 \ne \al_2 + \al_3$ has no solutions in conjugate algebraic numbers of degree $8$. However, the proof is much longer and it is not practical to give it here. For details, refer to \cite{DJ} again. \smallskip \noindent \emph{Remark 2 .} Another way to prove the non-existence of solutions of the equation $\al_1 = \al_2 + \al_3$ when $d=8$ in Theorem \ref{du1} is to use Theorem~\ref{viens}. Indeed, by mapping $\al_1$ to $\al_j$ for $j=2,3,\dots,8$, we obtain eight equalities of the form $$\al_j=\al_{k(j)}+\al_{l(j)}$$ with $j=1,2,\dots,8$ and $k(j)< l(j)$ lying in the set $\{1,2,\dots,8\} \setminus \{j\}$. At least one of these indices, say, $k(1)=2$ appears at least twice on the right hand side of these eight equalities, namely, $\al_1=\al_2+\al_3$ and $\al_j=\al_2+\al_l$, where $j \ne 1$ and so $l \ne 3$. Subtracting the first equality from the second we find that $\al_1-\al_j=\al_3-\al_l$. Thus, $\al_1+\al_l=\al_3+\al_j$. By Theorem~\ref{viens}, this is only possible when $d=4$, contrary to $d=8$. We must stress, however, that from the computational point of view, there is not much sense in proving Theorem \ref{du1} through Theorem \ref{viens}, since the direct search on the small list of candidates in Table \ref{pisot_d8} is much simpler than the large amount of computations required in the proof of Theorem \ref{viens}. \section{Beginning of the proof of Theorem~\ref{viens}}\label{proof_viens} Let $\mu, \nu \in \{-1, 1\}$. Assume that there is a Pisot number $\al$ whose four distinct conjugates $\al_1,\al_2,\al_3,\al_4$ satisfy the additive relation \[ \mu\al_1+\nu\al_2+\al_3+\al_4=0. \] The choice of parameters $\mu=\nu = -1$ yields the first equation of the theorem $\al_1 + \al_2 = \al_3 + \al_4$. The choices $\mu = -1, \nu = 1$ and $\mu = \nu = 1$ correspond to the equations $\al_1 = \al_2 + \al_3 +\al_4$ and $\al_1 + \al_2 + \al_3 + \al_4 = 0$, respectively. By applying an automorphism $\sigma$ of the Galois group $\gal(\Q(\al)/\Q)$ that sends $\al_1$ to $\al$, we deduce $\mu\al=-\nu\sigma(\al_2)-\sigma(\al_3)-\sigma(\al_4)$. Since $|\sigma(\al_j)|<1$ for $j=2,3,4$, this yields $\al<3$. This time we will apply Lemma~\ref{sammm} to \[ \be_1:=\mu\al_1, \quad \be_2:=\nu\al_2, \quad \be_3:=\al_3, \quad \be_4:=\al_4 \text{ and } N:=0. \] One needs to verify the condition \eqref{patikr}. Observe that the equation \[ \be_1^{-1}+\be_2^{-1}+\be_3^{-1}+\be_4^{-1}= \mu\al_1^{-1}+\nu\al_2^{-1}+\al_3^{-1}+\al_4^{-1}=0, \] yields \[ (\mu\al_1+\nu\al_2)\al_3\al_4=-\mu\nu(\al_3+\al_4)\al_1\al_2. \] Clearly, $\al_3+\al_4 \ne 0$, since otherwise $-\al$ is a conjugate of $\al$, which is impossible. Thus, dividing both sides by $-(\al_3+\al_4)=\mu\al_1+\nu\al_2 \ne 0$ we obtain the multiplicative relation $\al_3 \al_4=\mu\nu \al_1 \al_2$. Squaring both sides yields $\al_1^2\al_2^2 =\al_3^2\al_4^2$. In view of Lemma~\ref{mignot}, such an identity cannot hold, since the numbers $\al_1^2$, $\al_2^2$, $\al_3^2$ and $\al_4^2$ are distinct conjugates of a Pisot number $\al^2$. Therefore, by Lemma~\ref{sammm}, we obtain \[ 4h(\al)=h(\mu\al_1)+h(\nu\al_2)+h(\al_3)+h(\al_4) \geq \frac{1}{2} \log \left(\frac{1+\sqrt{5}}{2}\right). \] In view of $h(\pm\al)=\log{M(\al)}/d$ and $M(\al)=\al$, one can rewrite the last inequality as \begin{equation}\label{bound2} \frac{4\log{\al}}{d} \geq \frac{1}{2}\log{\left(\frac{1+\sqrt{5}}{2}\right)}. \end{equation} Since $1 < \al < 3$, the inequality \eqref{bound2} yields \begin{equation}\label{d18} d \le \frac{8 \log 3}{\log \left(\frac{1+\sqrt{5}}{2}\right)}=18.26409\dots. \end{equation} Consequently, the degree $d$ of $\al$ can only take the values in the range $4 \leq d \leq 18$. We can also rewrite inequalities \eqref{bound2}, \eqref{d18} as \begin{equation}\label{interv4} \left(\frac{1+\sqrt{5}}{2}\right)^{d/8} \le \al < 3, \quad 4 \leq d \leq 18. \end{equation} In the remainder of this section we will show that in case $d=4$ the only Pisot number $\al$ whose conjugates satisfy \eqref{m2} is precisely the number $(1+\sqrt{3+2\sqrt{5}})/2$. Assume that $\al$ is a Pisot number of degree $4$ with conjugates $\al=\al_1,\al_2,\al_3,\al_4$ and trace $t=\al_1+\al_2+\al_3+\al_4 \in \Z$. From $\al_1+\al_2=\al_3+\al_4$, we see that $\al_1+\al_2=t/2$. Next, from $\al_1+\al_2=\al+\al_2>0$ and $|\al_1+\al_2|=|\al_3+\al_4|<1+1=2$ we obtain $0<t/2<2$. Hence, $t \in \{1,2,3\}$. Furthermore, $t/2=\al_1+\al_2$ is the sum of two algebraic integers, so $t/2$ is an algebraic integer. Thus, the only choice for $t$ is $t=2$. Consequently, $\al_1+\al_2=\al_3+\al_4=1$. It follows that the minimal polynomial $f(x) \in \Z[x]$ of $\al$ is \begin{align*} f(x) &=(x-\al_1)(x-\al_2)(x-\al_3)(x-\al_4) =\\ &= (x^2-x+\al_1\al_2)(x^2-x+\al_3\al_4)=\\ &=(x^2-x+\be_1)(x^2-x+\be_2)=\\ &=(y+\be_1)(y+\be_2) =: g(y), \end{align*} where $y:=x^2-x$. Therefore, $\be_1=\al_1\al_2=\al_1(1-\al_1)<0$ and $\be_2=\al_3\al_4$ are real quadratic algebraic integers that are conjugate over $\Q$ (since $f(x)$ is irreducible). Since $\al_1 > 1$, $\al_2 = 1-\al_1$ and from the fact that $\al_1$ is a Pisot number, it follows that $\al_2 \in (-1, 0)$. From Lemma~\ref{duu} $(iii)$, it follows that $\be_1 \in (-2,0)$. Similarly, as $|\al_3|<1$ and $|\al_4|<1$, Lemma~\ref{duu} $(i)$ and $(ii)$ implies that $\be_2$ must be in the interval $(0,1)$. Thus, by Lemma~\ref{duu1}, the quadratic polynomial $g(-y)=(y-\be_1)(y-\be_2) \in \Z[y]$ must be $y^2+y-1$, which gives \[ \be_1 = \frac{-1-\sqrt{5}}{2}, \quad \be_2 = \frac{-1+\sqrt{5}}{2}. \] Hence, $g(y)=y^2-y-1$. By inserting $y=x^2-x$, we find that \[ f(x) =(x^2-x-\be_1)(x^2-x-\be_1)=g(x^2-x)=x^4-2x^3+x-1 \] with the root $\al=(1+\sqrt{3+2\sqrt{5}})/2$, as claimed. Clearly, the roots $\al_1,\al_2,\al_2,\al_4$ of $f(x)$ satisfy $\al_1+\al_2=\al_3+\al_4=1$. It remains to show that in cases $5 \le d \le 18$, no four distinct conjugates of a Pisot number of degree $d$ satisfy \eqref{m2}, and there exists no Pisot number $\al$ whose conjugates satisfy $\al_1 = \al_2 + \al_3 + \al_4$ or $\al_1 + \al_2 + \al_3 + \al_4 = 0$ for $4 \leq d \leq 18$. \section{Calculations and the end of the proof of Theorem \ref{viens}}\label{calc} Computationally, the problem is primarily one of finding the minimal polynomials of all Pisot numbers in subintervals of the interval $[1,3]$ of the appropriate degree, and then testing the equations \eqref{m2} and \eqref{lin3}. To find all Pisot numbers in a given interval up to a given degree, we wrote a fast implementation of the Boyd's algorithm \cite{Boyd78, Boyd84, Boyd85} in the {\tt C} language. Our implementation is based on {\tt FLINT} ({\bf F}ast {\bf Li}brary for {\bf N}umber {\bf T}heory) version 2.4.4 \cite{flint}. The {\tt FLINT} library provides the implementations of the polynomials in $\Z[x]$ and $\Q[x]$ with coefficients of arbitrarily large size and is highly optimized for fast arithmetical operations. The initial searches for small degrees (up to $d \leq 17$) were done on a single RedHat Linux server equipped with two Intel Xeon X5672 series 3.20GHz 12MB Cache 1333MHz 95W CPUs and 96735MB of RAM that was running at the University of Waterloo computing facilities. First we ran the search to find all the Pisot numbers up to degree $d \leq 8$ in the interval $[1, 2]$. The program found $109$ such Pisot numbers. We processed this list on {\tt Maple} and removed the numbers $\al \not\in (\tau^{d/6}, 2)$ that do not satisfy the inequalities \eqref{interv3}, leaving only $78$ Pisot numbers. Totals from the final list are recorded in Table \ref{countss1}. \begin{table}[h]\caption{\# of Pisot numbers $\al \in (\tau^{d/6}, 2)$, for $3 \leq d \leq 8$} \begin{tabular}{ccccccc} \toprule $\deg \al$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ \\ \midrule \# of $\al$'s & $4$ & $4$ & $12$ & $14$ & $24$ & $20$ \\ \bottomrule \end{tabular}\label{countss1} \end{table} The Pisot numbers for degrees $d=3, 6$ and $9$ satisfying $\alpha \in (\tau^{d/6},2)$ are given in Tables \ref{pisot_d3} \ref{pisot_d6} and \ref{pisot_d8}. These were manually checked in Section \ref{du_d7} to prove the remaining cases of Theorem \ref{du1}. To complete the proof of Theorem \ref{viens}, one needs to find all Pisot numbers up to degree $18$ in the interval $(1, 3)$ satisfying \eqref{interv4}. The searches on a single computer were feasible for $d \leq 17$, but for $d=18$, this was no longer practical: incrementing the degree $d$ by $1$ on the same interval resulted in multiplying the search time by a factor in the range $4$ to $4.4$, and doubling the memory usage. The predicted the search time on a single machine would be up to two CPU months while the RAM usage was predicted to remain under $1$ GB. The single machine search timings are recorded in Table \ref{singleCPU}. \begin{table}[h]\caption{Single Intel Xeon 3.4GHz machine search timings} \begin{tabular}{lll} \toprule $ \deg{\al}$ & Search interval & CPU time \\ \midrule $10$ & $[1, 3]$ & $25$ sec. \\ $11$ & $[1, 3]$ & $2$ \ min. $13$ sec.\\ $12$ & $[1, 3]$ & $11$ min. $7$ \\ $13$ & $[1, 3]$ & $54$ min. \\ $14$ & $[1, 3]$ & $4$ \ h. $13$ min. \\ $15$ & $[1, 3]$ & $18$ h. $47$ min. \\ $16$ & $\left[610/233, 3\right]$ & $3$ \days$11$ h. \\ $17$ & $\left[ 367/132, 3\right]$ & $13$ days $17$ h. \\ $18$ & $\left[437/148, 3\right]$ & \textcolor{Gray}{$\leq 60$ days (estimated)}\\ \bottomrule \end{tabular}\label{singleCPU} \end{table} Consequently, we decided to distribute the computations on a large collection of 2 Intel 5272 series 3.4Ghz/6M/1600Mhz 80W Dual Core Xeon Processor machines, allowing up to 120 simultaneous searches to be done. This was achieved by partitioning the search interval $[1,3]$ into 2868 subintervals. The lengths of these subintervals were balanced with the degree of the polynomials being searched. By inequality \eqref{interv4}, the closer a Pisot number $\al$ is to $x=3$ on the real line, the larger are the degrees that must be search for, and in turn, the smaller the subintervals that are searched. This is necessary because for any fixed interval there are considerably more Pisot numbers of large degree in this interval than Pisot numbers of smaller degrees. For example, our search found $40\,875$ Pisot numbers of degree 12 in the interval $[2, 3]$, while a much shorter interval between the points $ 1126/405 \approx 2.7802$ and $3$ contains $630\,165$ Pisot numbers of degree $17$. In practice, the intervals up to degree $d = 12$ were of fixed length $\eps=1/10$. For $d \geq 12$, we used the intervals of length $\eps(d) = 3^{12-d}\cdot 10^{-1}$. Distributed computations took 1$3.64$ CPU days. In total, $1\,956\,289$ Pisot numbers were found. It should be noted that the number of Pisot polynomials found is somewhat higher than actual number of Pisot numbers satisfying the inequality \eqref{interv4}. This is because we searched over a slightly larger collection of intervals $[a, b]$ with rational endpoints $a, b$ that cover all intervals $(\tau^{d/8}, 3)$, for $4 \leq d \leq 18$. After the minimal polynomials were computed, they were sieved by checking if their Pisot roots lie in intervals restricted by the inequalities \eqref{interv4}. All $1\,955\,183$ such Pisot numbers are counted in Table \ref{count222}. %, for each degree $d$ in range $4 \leq d \leq 18$. \begin{table}[h]\caption{\# of Pisot numbers$\al \in (\tau^{\deg{(\al)}/8}, 3)$} \begin{tabular}{llllll} \toprule $\deg{\al}$ & \# of$\al$'s & $\deg{\al}$ & \# of$\al$'s & $\deg{\al}$ & \# of$\al$'s \\ \midrule $4$ & $43$ & $9$ & $5\,555$ & $14$ & $140\,587$\\ $5$ & $162$ & $10$ & $9\,937$ & $15$ & $273\,851$\\ $6$ & $353$ & $11$ & $23\,410$ & $16$ & $402\,209$\\ $7$ & $1\,075$ & $12$ & $40\,812$ & $17$ & $630\,025$\\ $8$ & $2\,069$ & $13$ & $85\,979$ & $18$ & $339\,116$\\ \bottomrule \end{tabular}\label{count222} \end{table} By the result of Kurbatov stated as Lemma \ref{kurbat} in Section \ref{aux}, $d \not\in \{5,7,11,13,17\}$. So the next step of the sieve was to remove those polynomials of prime degree. This reduced the number of eligible Pisot polynomials to $1\,214\,532$. The third step was to find the minimal polynomials of Pisot numbers whose roots satisfy one of the three numerical inequalities \[ |\alpha_1 - \alpha_2 - \alpha_3 - \alpha_4 | < 10^{-5}, \qquad |\alpha_1 + \alpha_2 + \alpha_3 + \alpha_4 | < 10^{-5}, \] or \[ |\alpha_1 + \alpha_2 - \alpha_3 - \alpha_4 | < 10^{-5}. \] These calculations were done with 10 digits of accuracy, so this would get all potential Pisot numbers satisfying equation the equations of Theorem \ref{viens}. Some of these \emph{false positive} examples are shown in Table \ref{tab:fake}. \begin{table}[h]\caption{False positive examples, where $\alpha_1 + \alpha_2 \approx \alpha_2 + \alpha_4$} \begin{tabular}{p{3in}c} \toprule \\\\\\\\\\\\Pisot polynomial$f(x)$ & $|\alpha_1 + \alpha_2 - \alpha_3 - \alpha_4|$\\ \midrule $x^{15}-3 x^{14}+x^{13}+x^{12}-2 x^{11}+2 x^{10}-2 x^9+x^8+x^7-2 x^6+2 x^5-2 x^4+x^3+x^2-2 x+1 $ & $ 0.61690 \times 10^{-8} $\\ \\ $x^{15}-2 x^{14}-2 x^{13}-x^{12}-3 x^{11}-3 x^{10}-2 x^8-2 x^7-x^6-x^5-2 x^4-x^3-x^2-x-1 $ & $ 0.16262 \times 10^{-7} $\\ \\ $x^{18}-x^{17}-3 x^{16}-5 x^{15}-7 x^{14}-8 x^{13}-7 x^{12}-6 x^{11}-4 x^{10}-2 x^9+x^7+x^6+x^5-x^3-x^2-x-1 $ & $ 0.34922 \times 10^{-7} $\\ \\ $x^{18}-2 x^{17}-2 x^{16}-2 x^{15}-2 x^{14}-x^{13}-2 x^{12}-2 x^{11}-x^{10}-x^9+x^8+x^7+x^6+x^5+x^4+2 x^3+2 x^2+x+1 $ & $ 0.18618 \times 10^{-6} $\\ \\ $x^{16}-3 x^{15}+x^{14}-3 x^{13}+2 x^{12}-2 x^{11}+x^{10}-2 x^9-2 x^7+x^6-2 x^5+2 x^4-3 x^3+x^2-2 x+1 $ & $ 0.19425 \times 10^{-6} $\\ \\ $x^{16}-2 x^{15}-2 x^{14}-2 x^{13}-x^{11}-2 x^{10}-2 x^9-x^8-x^6-2 x^5-x^4-x-1 $ & $ 0.20095 \times 10^{-6} $\\ \\ $x^{16}-3 x^{15}+2 x^{13}-x^{12}-x^{11}+x^{10}+x^9-2 x^8+x^6-x^5+x^2-1 $ & $ 0.21102 \times 10^{-6} $\\ \\ $x^{16}-2 x^{15}-2 x^{14}-2 x^{13}-2 x^{12}-x^{11}-2 x^{10}-x^9+x^4-x-1 $ & $ 0.23696 \times 10^{-6} $\\ \\ $x^{15}-3 x^{14}+x^{11}-x^{10}+x^9-x^8+x^7+x^5-x^4-x^2+x-1 $ & $ 0.29620 \times 10^{-6}$\\ \bottomrule \end{tabular}\label{tab:fake} \end{table} In fact, this step resulted in a massive reduction of the list of candidates, with only $489$ Pisot numbers surviving: $271$ possible solutions to the equation $\al_1 + \al_2 = \al_3 + \al_4 $, $45$ possible solutions to the equation $\al_1 = \al_2 + \al_3 + \al_4$ and $173$ possible solutions to the equation $\al_1 + \al_2 + \al_3 + \al_4 = 0$. In the last step, for each of the remaining $489$ minimal polynomials of Pisot numbers (including degree $4$ polynomials), the resultant polynomials $g(x) = \Res_y(f(x-y), f(y))$ and $h(x) = \Res_y(f(x+y), f(y))$ were calculated. The resultant polynomials were tested by checking the conditions described in Lemma \ref{resultant}. In particular, the Pisot polynomials that pass the resultant test given in part $(iii)$ of Lemma \ref{resultant} must have four distinct roots $\al_1,\al_2,\al_3,\al_4$ satisfying $\al_1+\al_2=\al_3+\al_4$. As a result, the only example that was found to pass this test was the original example given in \cite{smydub0}, namely, the polynomial $ f(x)= x^4-2x^3+x-1. $ No polynomial passed the tests $(iv)$-$(v)$ of Lemma \ref{resultant}. Thus, the equations $\al_1 = \al_2 + \al_3 + \al_4$ and $\al_1 + \al_2 + \al_3 + \al_4 = 0$ cannot be solved in conjugates of any Pisot number. Therefore, the proof of Theorem \ref{viens} is completed. Finally, we remark that all the post-processing was done with {\tt Maple} on the Mac Book air x86\_64 machine. In total, it took$55.72$ CPU hours. \begin{thebibliography}{99} %\bibitem{bds} %{\sc G. Baron, M.~Drmota, M.~Skalba,} {\it Polynomial relations of %polynomial roots}, J. Algebra {\bf 177} (1995), 827--846. \bibitem{bert}{\sc M.J.~Bertin, A.~Decomps-Guilloux, M.~Grandet-Hugo, M.~Pathiaux-Delefosse and J.P.~Schreiber,} \emph{Pisot and Salem numbers,} Birkh\"auser, Basel, 1992.\bibitem{beu}{\sc F.~Beukers and D.~Zagier,} \emph{Lower bounds of heights of points on hypersurfaces,} Acta Arith. {\bf 79} (1997), 103--111. \bibitem{Boyd78}{\sc D.~W. Boyd}, \emph{Pisot and Salem numbers in intervals of the real line,} Math. Comp. {\bf 32} (144) (1978), 1244--1260. \bibitem{Boyd84}{\sc D.W.~Boyd,} \emph{Pisot numbers in the neighborhood of a limit point. {II},} Math. Comp. {\bf 43} (168) (1984), 593--602. \bibitem{Boyd85}{\sc D.W.~Boyd}, \emph{Pisot numbers in the neighbourhood of a limit point. {I},} J. Number Theory {\bf 21} (1985), 17--43. \bibitem{dub}{\sc A.~Dubickas,}{\it On the degree of a linear form in conjugates of an algebraic number,} Illinois J. Math. {\bf 46} (2002), 571--585. %\bibitem{du0} %{\sc A.~Dubickas,} {\it Sumsets of Pisot and Salem numbers,} %Expo. Math. {\bf 26} (2008), 85--91. \bibitem{DJ}{\sc A.~Dubickas and J.~Jankauskas,} \emph{Simple linear relations with algebraic numbers of small degree,} submitted for publication, 2014. \bibitem{smydub0}{\sc A.~Dubickas and C.J.~Smyth,} \emph{On the lines passing through two conjugates of a Salem number,} Math. Proc. Camb. Phil. Soc. {\bf 144} (2008), 29--37. %\bibitem{ds1} %{\sc M.~Drmota, M.~Skalba,} {\it On multiplicative and linear %independence of polynomial roots}, in: Contributions to General Algebra 7 %(eds. D. Dorninger, G. Eigenthaler, H.K. Kaiser, and W.B. Muwller), %Hoelder--Pichler--Tempsky, Wien; Teubner, Stuttgart, 1991, 127--135. %\bibitem{ds2} %{\sc M.~Drmota, M.~Skalba,} {\it Relations between polynomial roots}, %Acta Arith. {\bf 71} (1995), 65--77. \bibitem{flint}{\sc W.~Hart, F.~Johansson and S.~Pancratz,} \emph{{FLINT}: Fast Library for Number Theory,} 2013, Version 2.4.4, \url{http://flintlib.org}. \bibitem{garza}{\sc J.~Garza, M.I.M.~Ishak and C.~Pinner,} \emph{On the product of heights of algebraic numbers summing to real numbers,} Acta Arith. {\bf 142} (2010), 51--58. \bibitem{schi1}{\sc G.~Hoehn and N.-P.~Skoruppa,} \emph{Un r\'esultat de Schinzel,} J. Th\'eor. des Nombres de Bordeaux{\bf 5} (1993), 185. \bibitem{kurb}{\sc V.A.~Kurbatov,} \emph{Galois extensions of prime degree and their primitive elements,} Soviet Math. (Izv. VUZ) {\bf 21} (1977), 49--52. \bibitem{la1}{\sc J.C.~Lagarias, H.A.~Porta and K.B.~Stolarsky,} \emph{Asymmetric tent map expansions I. Eventually periodic points,} J. London Math. Soc. (2) {\bf 47} (1993), 542--556. \bibitem{la2}{\sc J.C.~Lagarias, H.A.~Porta and K.B.~Stolarsky,} \emph{Asymmetric tent map expansions II. Purely periodic points,} Illinois J. Math. {\bf 38} (1994), 574--588. \bibitem{mig}{\sc M.~Mignotte,} \emph{Sur les conjugu\'es des nombres de Pisot,} C. R. Acad. Sci. Paris S\'er. I. Math.{\bf 298} (1984), 21. \bibitem{nar}{\sc W.~Narkiewicz,} \emph{Elementary and analytic theory of algebraic numbers,} 3rd ed., Springer, Berlin, Heidelberg, 2004. \bibitem{pad}{\sc R.~Padovan,} \emph{Dom Hans Van Der Laan and the plastic number}, Nexus IV: Architecture and Mathematics, Kim Williams Books, 2002, pp. 181--193. \bibitem{salem}{\sc R.~Salem,} \emph{A remarkable class of algebraic numbers. Proof of a conjecture of Vijayaraghavan,} Duke Math. J. {\bf 11} (1944), 103--108. \bibitem{sam}{\sc C.L.~Samuels,} \emph{Lower bounds on the projective heights of algebraic points,} Acta Arith. {\bf 125} (2006), 41--50. \bibitem{schi}{\sc A.~Schinzel,} \emph{On the product of the conjugates outside the unit circle of an algebraic integer,} Acta Arith. {\bf 24} (1973), 385--399. \bibitem{sieg}{\sc C.L.~Siegel,} \emph{Algebraic numbers whose conjugates lie in the unit circle}, Duke Math. J. {\bf 11} (1944), 597--602. \bibitem{smy0}{\sc C.~J.~Smyth,} \emph{On the product of the conjugates outside the unit circle of an algebraic integer,} Bull. London Math. Soc. {\bf 3} (1971), 169--175. \bibitem{smy1}{\sc C.J.~Smyth,} \emph{The conjugates of algebraic integers,} Amer. Math. Monthly {\bf 82} (1975), 86. \bibitem{my}{\sc C.J.~Smyth,} \emph{Conjugate algebraic numbers on conics,} Acta Arith. {\bf 40} (1982), 333--346. \bibitem{smy2}{\sc C.J.~Smyth,} \emph{There are only eleven special Pisot numbers,} Bull. London Math. Soc. {\bf 31} (1999), 1--5. \bibitem{smy3}{\sc C.J.~Smyth,} \emph{Mahler measure of one-variable polynomials: a survey.} Conference proceedings, University of Bristol, 3--7 April 2006, LMS Lecture Note Series 352, Cambridge University Press, Cambridge, 2008, pp. 322--349. \bibitem{smy4}{\sc C.J.~Smyth,} \emph{Seventy years of Salem numbers: a survey,} preprint, 2014, \break arXiv:1408.0195v2. \end{thebibliography} \end{document} }

179 — 1410.1647

\caption{ H~I column densities. \opensquare: Components~1. \fullsquare: targets presenting evidence for an astrosphere and only one H~I component, i.e. sight lines completely filled with the Local Cloud. Crossed \opensquare: targets where the presence of the astrosphere is uncertain. Diagonal lines show lines of constant densities $n({\rm H~I})=0.03,\,0.04,\,0.1\,{\rm cm}^{-3}$ . \label{fig:distanghi} }

180 — 1410.1841

\caption{\label{tab:models} Gravitational (GR) and non-gravitational (NG) solutions based on the data interval of 2012 October~4--2014 March~3.In the \textcolor{magenta}{\bf third NG~model}, parameters $A_1$ and $A_2$ was assumed according to Farnocchia solution given at \citet{JPL_Browser} at the end of September; this NG~determination was based on 180 measurements selected by Farnocchia from more than 2000 observations. }

\caption{\label{fig:New-old}\textcolor{red}{\footnotesize The $1/a_{\textrm{ori}}$ distribution for the observed sample of large perihelion Oort spike comets $q_{\textrm{osc}}>3$ AU (64 objects), studied in \citet{dyb-kroli:2011}. The filled part of histogram represents dynamically old comets, and a gray part - seven comets with unclear dynamical status. The uncertainty of the determination of $1/a_{\textrm{ori}}$ is taken into account for each comet.}}

181 — 1410.2795

\caption{(Color online) (a) Experimentally measured threshold indentation for wrinkling, $\woc$, as a function of $\lc(\glv/Y)^{1/2}$. Experiments with varying film thickness ($59\mathrm{~nm}\leq t\leq 246\mathrm{~nm}$), $\glv=72\mathrm{~mN/m}$ are shown for indenter radii $\rtip=135\mathrm{~\mu m}$ ({\color{green} $\blacktriangle$}) and $\rtip=25\mathrm{~\mu m}$ ({\color{red}{\large $\bullet$}}). Experiments with varying surface tension coefficient ($36\mathrm{~mN/m}\leq\glv\leq72\mathrm{~mN/m}$) and $t=121\mathrm{~nm}$ ({\color{blue} $\blacksquare$}). The theoretical prediction for $\Rnd\gg1$, $\tdelta_c\approx11.75$, is also shown (dashed line). Good agreement with experiment justifies our neglect of indenter size and any (hypothesized) manufacture-dependent pre-stress, which were both attributed crucial roles previously \cite{bernal11}. (b) The inner wrinkle radius, $r=\Li$, decreases with increasing indentation, $\tdelta$, when wrinkles reach the film's edge. Experiments with $\glv=72\mathrm{~mN/m}$ and: $t=85\mathrm{~nm}$ ($\red{\Box}$), $t=121\mathrm{~nm}$ ($\red{\bigcirc}$), $t=158\mathrm{~nm}$ ($\red{\bigtriangleup}$), $t=207\mathrm{~nm}$ ($\red{\times}$), $t=246\mathrm{~nm}$ (\red{\ding{73}}). Experiments with $t=121\mathrm{~nm}$ and: $\glv=58\mathrm{~mN/m}$ (${\color{green} \blacklozenge}$), $\glv=50\mathrm{~mN/m}$ (${\color{blue} \blacktriangleright}$) and $\glv=42\mathrm{~mN/m}$ ($ \blacktriangleleft$). The prediction of the FT theory (solid curve) and the asymptotic result \eqref{eqn:smalllamasy} (dashed line) are also shown. The wrinkle number scales similarly to that found in other studies \cite{huang07} (data not shown).}

182 — 1410.2948

\caption{(\textit{Color online}) DC susceptibility curves of Li$_{x}$RhB$_{1.5}$ (x=0.8, 1.0, 1.2) at 20 Oe on ZFC (open symbols) and\FC (filled symbols) cycles.\\textit{Inset}: an expanded view showing the typical character of the PME:\on FC, just below$% T_{c}$, $\protect\chi (T)$ becomes negative and afterwards, on further cooling, turns into a positive value. By contrast,\the ZFC susceptibility exhibits the normally-expected (negative) screening signal. The magnitude of the PME signal at$H$ = 20 Oe is $\sim $0.1\% of the shielding ZFC signal (this is reminiscent of the PME in Nb disk\protect\cite% {Thompson95-ParamMeissner-Nb}). The structure in both ZFC and FC curves within the immediate neighborhood below $T_{c}$ is related to the fact that, within this region, the critical currents associated with\most of the PME loops are too small to drive spontaneous moments.\protect\cite% {Khomskii94-ParamagMeissner,Li03-ParamagMeissner,Siegrist-ParamagMeissner-dwave}}

\caption{(\textit{Color Online}) (a, b, c):\Measured$H_{c1}(T)$ (symbols) of Li$_{x}$RhB$_{1.5}$ were fitted (solid lines) to the relation $H_{c1}=H_{c1}(0)(1-t^{2})$, where $t$ =$T/T_{c}$. (d, e, f): measured $H_{c2}(T)$ of Li$_{x}$RhB$_{1.5}$ (symbol) were analyzed using (i) the quadratic formula $H_{c2}$($t$) = $H_{c2}$[(1-$t^{2}$% )/(1+$t^{2}$)] (solid line) and (ii) the WHH expression (dashed lines).}

183 — 1410.3444

\caption{Low energy spectrum for $s=\frac{5}{2}$ as function of total spin $S$ with respect to the lowest energy of each $S$ sector $E_0(S)$. Each level belongs to an irreducible representation according to the key (the number in the parenthesis is the spatial degeneracy): $\circ$: $A_g$ (1), {\color{red}{$\Box$}}: $T_{1g}$ (3), {\color{green}{$\diamond$}}: $T_{2g}$ (3), {\color{blue}{$\vartriangle$}}: $F_g$ (4), {\color{brown}{$\triangleleft$}}: $H_g$ (5), {\color{violet}{$\triangledown$}}: $A_u$ (1), {\color{cyan}{$\triangleright$}}: $T_{1u}$ (3), {\color{magenta}{+}}: $T_{2u}$ (3), {\color{orange}{$\times$}}: $F_u$ (4), {\color{indigo(dye)}{$\ast$}}: $H_u$ (5). The spatial degeneracy has to be multiplied with the $2S+1$ degeneracy of each $S$ multiplet to give the total number of states corresponding to each symbol. Only a certain number of levels has been calculated for each $S$, which explains the lack of more excited states for intermediate $S$. %(/home/user1/basic/diag/fullerenes/run17) }

184 — 1410.4355

\caption{Ten most anomalous conferences for each year displayed with $p-$value and number of membership changes. Blue entries are \textcolor{BlueViolet}{true-positives} while red entries are \textcolor{BrickRed}{false-positives} with threshold $\alpha = 10^{-4}$.}

185 — 1410.4613

\caption{The network structure for the mechanical system. The nodes \textbf{G1} and \textbf{G2} denote the subsystems. The internal inputs (the inputs of each subsystem) are numbered sequentially in {\color{red}red} color. The internal outputs (the outputs of each subsystem) are numbered sequentially in {\color{cyan}cyan} color. The external inputs (the inputs of the overall interconnected system) are numbered sequentially in {\color{blue}blue} color. Finally, the external outputs (the outputs of the overall interconnected system) are numbered sequentially in {\color{green}green} color. The solid curves portray the edges between internal inputs, internal outputs, external inputs, and external outputs. The weights on the edges are displayed only if they are not equal to identity. Notice that if more than one edge is going to an input, implicitly, we mean that these edges are summed together before being fed to that input. \label{fig:matlab:networkStructure}}

186 — 1410.4953

\caption{Four different behaviours in the $(\tau, \omega)$ parameter space and the theoretical bifurcation curves for $\alpha = 0.01$, $N=200$, $\gamma=1$. The horizontal line represents the boundary of the parameter domain where the graph is connected. In the simulation, networks which on average had at least 3 disjointed components were considered disconnected. The other two curves are the transcritical bifurcation curves obtained from the mean-field approximation (continuous diagonal line) and from the pairwise approximation \eqref{tauTC} (dashed curve). The markers are as follows: \textcolor[rgb]{0.00,0.00,1.00}{$\times$} - connected, epidemic, \textcolor[rgb]{1.00,0.00,0.00}{$\times$} - connected, no epidemic, \textcolor[rgb]{0.00,0.00,1.00}{$\circ$} - disconnected, epidemic, and \textcolor[rgb]{1.00,0.00,0.00}{$\circ$} - disconnected, no epidemic.}

187 — 1410.5926

\caption{AUC: area under ROC curve (larger is better). The best three results are highlighted with {\color{red}red}, {\color{green}green}, and {\color{blue}blue} fonts, respectively.}

188 — 1410.5946

\caption{Graphs of {\color{red} $S(n_{\sigma};1.2\times10^{11},86)$ } (red) and {\color{blue} $S(n_{\sigma};1.6\times10^{11},96)$ } (blue). The smear is seen to be $\approx 3\%$ at 1.5 $\sigma$.}

189 — 1410.6001

\caption{\normalsize Overall performances ``$avgDCG$'' for several different time periods. ``$\uparrow$'' indicates the larger the value the better the performance. The bold number indicates the best performance. ``$\ast$'' indicates the best performance of multi-CG among the three indicators. The {\color{red}{\sl red italic}} number indicates the best performance of all the latent factors from Twitter for the indicator of trading volume. The {\color{blue}{\sl blue italic}} number indicates the best performance of all the latent factors from Twitter for the price related indicators.}

\caption{Top 4 correlations of trading volume (TV) for company APC. The company in {\color{red}{red}} indicates the real-world trading volume correlation for a certain time period. The company in bold indicates its appearance in the ground truth (i.e., the TV column). The company with ``$\ast$'' indicates its right rank discovered from Twitter.}

190 — 1410.6064

\caption{\red{Schematic representation showing the constituents of our biomolecular control system. The network on the right (inside the cloud) represents the open-loop network, whose dynamics are to be controlled. Control is achieved by augmenting another network of reactions, referred to as the controller network (outside the cloud). Together the two networks form the closed-loop network whose dynamics are determined by the coupling resulting from the interaction of both networks. The proposed controller network, which we refer to as antithetic integral controller, acts on the open-loop network by influencing the rate of production of the actuated species $\X{1}$ by means of the control input species $\Z{1}$. The regulated species $\X{\ell}$ will be influenced by the increase or decrease of the actuated species $\X{1}$ and, in return, will influence the rate of production of the sensing species $\Z{2}$, that will, finally, annihilate with the control input species $\Z{1}$, thereby implementing a negative feedback control loop. The integral action is encoded in all the reactions of the controller network.}}

191 — 1410.6544

\caption{VO87 diagnostic diagram log([N II]$\lambda$6583/H$\alpha$)$\sim$log([O III]$\lambda$5007/H$\beta$) when using log($t$/yr)=6.3 (young, left panel) and 7.5 (IA, right panel) SPs with (red, solid line) and without (black, dashed line) BIs in the case of $n_{\rm H}$=100${\rm cm^{-3}}$. % On each grid, symbols (+, $\ast$, $\circ$, $\times$, $\Box$, $\triangle$, $\oplus$) are for $Z$=0.0001, 0.0003, 0.001, 0.004, 0.01, 0.01 and 0.03 and line width increases with log$U$ (=$-$4,$-$3,$-$2, $-$1 and 0, from bottom to top). % Comparisons include the border lines (Dop00, Kew01, Kau03, Sta06; --- , -\,-\,-, -$\cdot$-$\cdot$, $\cdots$; grey) and observations (B05 [$\lozenge$], V98 [{\color{black} $\blacktriangle$}], S13[{\color{green} $\blacksquare$}], S10[{\color{blue} $\bullet$}], B09[{\color{cyan} $\bigstar$}], R08[{\color{magenta} $\square$}], W11[{\color{yellow} $\circ$}], V06[{\color{red} $\bigcirc$}]). % The absence of V89, R08, W11 and V06 observations in this plot is caused by the lack of corresponding data.}

192 — 1410.7038

\caption{%\begin{counted} \textbf{Phase diagram for spherical shells} Phase diagram for snapping behavior of spherical shells over a wide range of geometrical parameters $\alpha$ and $1/\sqrt{\gamma}$. Stability behavior in experiments is characterized as bi-stable (\textcolor{OliveGreen}{$\blacksquare$}, switches to folded state through a snapping mechanism), mono-stable (\textcolor{Crimson}{$\blacksquare$}, prefers unfolded state) or temporarily-stable (\textcolor{Navy}{$\blacksquare$}, closer to phase boundary: snap back on a time-scale of seconds without external perturbations). Finite element simulations (points solved denoted with \textbf{+}) provide regions of mono-stability (red shading) and bi-stability (green shading) and the theoretical scaling law agrees well with both experiments and simulations. Each experimental data point was analyzed for at least 3 shells of appropriate parameters. %\end{counted} \tcb{Caption has \thewords\ words} }

193 — 1410.7429

\caption{SR ($\times 3$) result comparison between the MMSE and MAP estimate (\textbf{PSNR/$100 \times$SSIM}). Better results of the first two rows (with the same FoE prior) are colored with {\color{blue}blue}. The best results are highlighted in bold.}

\caption{Noisy image SR ($\times 3$) result comparison between the MMSE and MAP estimate (\textbf{PSNR/$100 \times$SSIM}). Better results of the first two rows (with the same FoE prior) are colored with {\color{blue}blue}. The best results are highlighted in bold.}

194 — 1410.7587

\caption{ %Fig.2. The dependence of the relative critical current $I_{c}(P)/I_{c}(0)$ in the sample Sn1 on the reduced microwave irradiation power $P/P_{c}$ at $T= 3.812$~K for different irradiation frequencies $f$, GHz: 15.4~($\blacktriangle$), 8.1~(\fullcircle), 3.7~(\fullsquare) ($I_{c}(0)$ is the critical current of the film at $P=0$; $P_{c}$ is the minimum power of electromagnetic irradiation at which $I_{c}(P)=0$). }

\caption{\label{f3} The reduced value of exceeding the maximum critical current $I_{cmax}( P)$ over $I_{c}(0)$ as a function of the irradiation frequency for the samples Sn1~($\blacktriangle$), SnW10~(\fullsquare) and SnW5~(\fullcircle) at $T/T_{c}\approx$ 0.99; the values of lower cut-off frequencies of superconductivity enhancement calculated by \Eref{Eliashberg11} for the samples Sn1~(\opentriangle), SnW10~(\opensquare)) and SnW5~(\opencircle).}

\caption{\label{f4} The dependence of the relative critical current $I_{c}(P)/I_{c}(0)$ in the sample SnW10 on the reduced microwave irradiation power $P/P_{c}$ at $T= 3.777$~K for different irradiation frequencies $f$, GHz: 12.91~(\fullsquare), 6.15~(\fullcircle), 0.63~($\blacktriangledown$); dashed curve 3 is the dependence $I_{c}(P)/I_{c}(0)(P/P_{c})$ calculated by \Eref{Eliashberg21}.}

\caption{\label{f5} The dependence of the relative critical current $I_{c}(P)/I_{c}(0)$ in the sample SnW5 on the reduced microwave radiation power $P/P_{c}$ at $T = 3.744$~K for different irradiation frequencies $f$, GHz: 15.2 ($\blacktriangle$), 11.9~(\fullcircle), 9.2~(\fullsquare), 5.6~($\blacktriangledown$). The dashed curve 2 is the dependence ($I_{c}(P)/I_{c}(0))(P/P_{c}$) calculated by \Eref{Eliashberg21}.}

\caption{ %Fig.6. The dependence of the relative critical current $I_{c}(P)/I_{c}(0)$ on the reduced microwave irradiation power $P/P_{c}$ with the frequency $f$= 9.2~GHz at $T/T_{c} \approx$ 0.99 in different samples: SnW5~($\blacktriangle$), SnW6~(\fullcircle) and SnW10~(\fullsquare). }

\caption{ %Fig.8. The dependence of the critical current $I_{c}$ (\fullsquare) and the maximum current of existence of the vortex resistivity $I_{m}$ (\fullcircle) for the sample SnW5 on the reduced microwave power $P/P_{c}$ with the frequency $f$=12.89~GHz at $T = 3.748$~K. The inset shows an enlarged fragment of the above-mentioned dependencies. }

\caption{ %Fig.9. The experimental temperature dependence of the critical currents $I_{c}(P=0)$~(\fullsquare), $I_{c}(f = 9.2~GHz)$~(\fullcircle), and $I_{c}(f = 12.9~GHz)$~($\blacktriangledown$) for the sample SnW10. The theoretical dependence $I_{c}^{GL}(T)=7.07\times 10^{2}(1 - T/T_{c})^{3/2}$~mA calculated by \Eref{Eliashberg19} \cite{DmitrievZolochevskiiSalenkovai25} (curve 1); calculated dependence $I_{c}(T)=5.9\times 10^{2}(1 - T/T_{c})^{3/2}$~mA (curve 2); theoretical dependence $I_{c}^{AL}(T) =9.12\times 10^{1}(1 - T/T_{c})$~mA calculated by \Eref{Eliashberg27} \cite{AslamazovLempitsky}~(straight line 3); theoretical dependence $I_{c}(f = 9.2~GHz)$ calculated by \Eref{Eliashberg16} and fitting dependence $I_{c}(T)=6.5\times 10^{2}(1 - T/3.818)^{3/2}$~mA (curve 4); theoretical dependence $I_{c}(f=12.9~GHz)$ calculated by \Eref{Eliashberg16}, and fitting dependence $I_{c}(T)=6.7\times 10^{2} (1 - T/3.822)^{3/2}$~mA (curve 5); theoretical dependence $I_{c}(f = 9.2~GHz)$ calculated by \Eref{Eliashberg16} normalized by the curve 2, and fitting dependence $I_{c}(T)=5.9\times 10^{2}(1 - T/3.818)^{3/2}$~mA (curve 6); calculated dependence $I_{c}( T)=9.4\times 10^{1}(1 - T/3.818)$~mA (straight line 7). }

\caption{ \label{f11} %Fig.11. The experimental temperature dependencies of the maximum current $I_{m}$ of existence of stationary uniform flow of intrinsic vortices of transport current across the film SnW5: $I_{m}(T,P = 0)$ (\fullcircle), $I_{m}^{P}(T,f=9.2~GHz)$~($\blacktriangledown$) and $I_{m}^{P}(T,f=15.2~GHz$)~($\blacktriangle$). Curve 1 is the theoretical dependence $I_{m}(T)$ (see \Eref{Eliashberg23}); curve 2 is the calculated dependence $I_{m}^{P}(T,f=9.2~GHz$) (see \Eref{Eliashberg24}); curve 3 is the calculated dependence $I_{m}^{P}(T,f=15.2~GHz$) (see \Eref{Eliashberg26}).}

195 — 1410.7628

\caption{\label{fig:fig3} Potential of the model 1D crystal after applying Gaussian-type inhomogeneous strain. $V(x') = U^*(x;\varepsilon(x)) = -\left[U_0+K\varepsilon(x)\right] \cos(4\pi x/a_0)$. $x$ is related to $x'$ via $x' = x + \frac{\sqrt{\pi} L}{8}\varepsilon_{\rm max} \left[\mathrm{erf} \left(\frac{x-L/2}{L/4}\right) - \mathrm{erf}(-2) \right]$. The values of model parameters $U_0$, $K$, $L$ and $\varepsilon_{\rm max}$ are given by $U_0 = 0.2$, $K = -0.5$, $ L = 25a_0$ and $\varepsilon_{\rm max} = 0.1$.} \end{figure} %%%%% With the above model set-up, we now apply a Gaussian-type inhomogeneous strain on the 1D crystal. The strain distribution is given by %%%%% \begin{equation} \varepsilon(x) = \varepsilon_{\rm max} \exp\left[-\frac{(x-L/2)^2}{(L/4)^2}\right], \label{eq:Inhomo_strain_field_1D} \end{equation} %%%%% where $\varepsilon_{\rm max}$ is the maximum strain value in the strain field $\varepsilon(x)$, occurring at $x = L/2$. $L$ denotes the size of crystal. After applying the inhomogeneous strain, a position $x$ in the undeformed crystal will map to a new position $x'$ in deformed coordinates given by %%%%% \begin{equation} \begin{split} x' & = x + \int_0^x \varepsilon(v) dv \\ & = x + \frac{\sqrt{\pi} L}{8}\varepsilon_{\rm max} \left[\mathrm{erf} \left(\frac{x-L/2}{L/4}\right) - \mathrm{erf}(-2) \right], \end{split} \end{equation} %%%%% where $\mathrm{erf}(x)$ denotes error function. Denote by $V(x')$ the crystal potential of the 1D crystal after applying the Gaussian inhomogeneous strain, we then adopt the local approximation of crystal potential as described in Sec.~\ref {subsec:locality}, which says that the inhomogeneously strained crystal potential at point $x'$ can be well approximated by the crystal potential of a homogeneously strained crystal with the same strain value. This can be mathematically written out as %%%%% \begin{equation} V(x') \approx U^*(x; \varepsilon(x)), \quad x' = x + \int_0^x \varepsilon(v) dv. \end{equation} %%%%% The as-constructed strained crystal potential $V(x')$ is visualized in Fig.~\ref{fig:fig3}. We have thus, for demonstration purpose, explicitly constructed the strained crystal potential $V(x')$ using the local approximation of crystal potential. This allows us to solve the energy eigenstates of an inhomogeneously strained crystal using two distinct methods: Method 1: direct numerical diagonalization of strained Hamiltonian. Since the explicit expression for the inhomogeneously strained crystal potential $V(x')$ has been constructed, we can solve the Schr\"{o}dinger equation for the inhomogeneously strained crystal in deformed coordinates, %%%%% \begin{equation} \left[-\frac{1}{2}\frac{d^2}{dx'^2} + V(x') \right] \Psi(x') = E \Psi(x'), \label{eq:1D_Deformed} \end{equation} %%%%% by diagonalizing the Hamiltonian $H = -\frac{1}{2}\frac{d^2}{dx'^2} + V(x')$ using plane wave basis set in Fourier space. More straightforwardly, we can discretize the wavefunction $\Psi(x')$ into a $N \times 1$ matrix vector in real space, %%%%% \begin{equation} \Psi(x') = \begin{bmatrix} \Psi(x'_1) & \Psi(x'_2) & \cdots & \Psi(x'_N) \end{bmatrix}^T, \end{equation} %%%%% and then write the Hamiltonian as a matrix operator $\mathbf{H}$ acting on the wavefunction $\mathbf{H} = -\mathbf{L}/2 + \mathbf{V}$, where $\mathbf{L}$ and $\mathbf{V}$ are the matrix operators for the differential operator $\frac{d^2}{dx'^2}$ and the potential operator $V(x')$ respectively: \begin{equation} \mathbf{L} = \frac{1}{(\Delta x')^2} \begin{bmatrix} -2 & 1 & & 1 \\ 1 & -2 & 1 & \\ & 1 & \ddots & 1 \\ 1 & & 1 & -2 \end{bmatrix}, \end{equation} \begin{equation} \mathbf{V} = \begin{bmatrix} V(x'_1)& & & \\ & V(x'_2) & & \\ & & \ddots & \\ & & & V(x'_N) \end{bmatrix}. \end{equation} %%%%% $\Delta x' = x'_{i+1} - x'_{i}$ is the distance between two real space grid points. The Hamiltonian matrix $\mathbf{H}$ can then be numerically diagonalized to obtain the energy eigenvalues $E$ and wavefunctions $\Psi(x')$. Method 2: solving the energy eigenstates of inhomogeneously strained crystal using our envelope function method. We can solve the Schr\"{o}dinger equation, Eq.~\ref{eq:1D_Deformed}, by first mapping it back to undeformed coordinates, which becomes %%%%% \begin{equation} \left[\mathcal{P}^* + U^*(x;\varepsilon(x))\right] \Psi^*(x) = E\Psi^*(x). \label{eq:1D_undeformed_coordinates} \end{equation} %%%%% The explicit expression for the differential operator $\mathcal{P}^*$ is given by Eq.~\ref{eq:OperatorP_1D}. The mapped wavefunctions $\Psi^*(x)$ will then be expressed in terms of envelope functions $F_n(x)$ and strain-parametrized Bloch functions $\ux$: %%%%% \begin{equation} \Psi^*(x) = \sum_n F_n(x) \ux. \label{eq:1D_envelope_expansion} \end{equation} %%%%% We then follow the procedures described in Sec.~\ref{sec:1DFramework} to eliminate the crystal potential term $U^*(x;\varepsilon(x))$ in Eq.~\ref{eq:1D_undeformed_coordinates} using strain-parametrized Bloch functions $\ux$ and the associated strain-parametrized energy eigenvalues $\epsilon(x; \varepsilon(x))$. Eq.~\ref{eq:1D_undeformed_coordinates} can then be turned into a coupled differential eigenvalue equation for the envelope functions $F_n(x)$ given by Eq.~\ref{eq:Envelope_Function_1D}, and solved as a generalized matrix eigenvector problem. The solution of Eq.~\ref{eq:Envelope_Function_1D} requires the explicit construction of strain-parametrized functions $\ux$ and the associated strain-parametrized energy eigenvalues $\epsilon(x; \varepsilon(x))$. The construction of these functions involves unit-cell level calculations of homogeneously strained 1D crystals. Only the Bloch functions and energy eigenvalues of the electronic states at the reference crystal momentum ($k = 0$ in this case) and a few bands close to the valence/conduction band need to be calculated. The homogeneous strain values $\varepsilon$ are coarsely taken from the inhomogeneous strain field (no more than one grid point per unit cell). The calculated periodic Bloch functions of each homogeneously strained crystal are then expressed in plane wave basis as $u_{n0}^*(x;\varepsilon) = \sum_m C_m^n(\varepsilon) e^{i2\pi mx/a_0}$, where $C_m^n(\varepsilon)$ are the expansion coefficients. The strain-parametrized Bloch functions can then be constructed by letting $\varepsilon = \varepsilon(x)$ at position $x$, namely %%%%% \begin{equation} u_{n0}^*(x; \varepsilon(x)) = \sum_m C_m^n(\varepsilon(x)) e^{i2\pi mx/a_0}. \end{equation} %%%%% It is easy to see that, at position $x$, the value of $u_{n0}^*(x; \varepsilon(x))$ is the same as the value of periodic Bloch function $u_{n0}^*(x;\varepsilon)$ at $x$ and $\varepsilon = \varepsilon(x)$. It is in this sense $u_{n0}^*(x; \varepsilon(x))$ are named strain-parametrized functions. Since $C_m^n(\varepsilon)$ are smooth functions of $\varepsilon$ and $\varepsilon(x)$ is a smooth function of $x$, we can use polynomial fitting to obtain smooth functions for $C_m^n(\varepsilon(x))$. As only unit-cell level calculations of homogeneously strained crystals at a reference crystal momentum are involved, the construction of the strain-parametrized functions $\ux$ and $\epsilon(x; \varepsilon(x))$ do not require much computational power in this 1D example. \begin{figure}[t] \includegraphics[width=0.45\textwidth]{fig4} \caption{\label{fig:fig4} (a) Energy eigenvalues of the unstrained (open circles) and inhomogeneously strained model 1D crystal (filled circles) obtained by direct diagonalization. The energy levels are shifted horizontally with respect to each other to resolve energy levels which are very close to each other. The energy range of hole and electron bound states in strained crystals are labeled. (b) Wavefunction probability amplitude for hole and electron bound states, which are labeled in the figure as VBM, VBM-1, CBM, and CBM+1. VBM denotes valence band maximum; VBM-1 denotes one energy level below VBM; CBM means conduction band minimum, while CBM+1 denotes one energy level above CBM. The wavefunctions have rapid oscillation.} \end{figure} Of the two methods discussed above, Method 1, the direct diagonalization of Hamiltonian, is a well established method, therefore it can be used to benchmark Method 2, our envelope function method. To test the effectiveness of our envelope function method, we have calculated the energy eigenvalues and eigenfunctions of the 1D inhomogeneously strained crystal using both methods. A special note is that we are not testing here how good the local approximation of crystal potential for inhomogeneously strained crystal can be, but how accurate and fast our envelope function method can achieve given the local approximation of crystal potential is a sufficiently good approximation. Also note that, although for the sake of benchmarking our envelope function method, we have explicitly constructed the strained crystal potential in this 1D problem, in practical application of our envelope function method, such explicit construction of crystal potential will not be performed. The information of local strained crystal potential, at the level of approximation used in our method, is reflected in the strain-parametrized Bloch functions $\ux$ and the associated strain-parametrized energy eigenvalues $\epsilon_{n0}(x;\varepsilon(x))$. Choosing the following model parameters $L = 100a_0$, $U_0 =0.2$, $K = -0.5$, and $\varepsilon_{\rm max} = 0.1$ for the 1D inhomogeneously strained crystal, we carry out numerical real space diagonalization of the Hamiltonian by spatially discretizing the wavefunction $\Psi(x)$ into a $N\times 1$ matrix. Periodic boundary condition $\Psi(0) = \Psi(L)$ is adopted. As the wavefunction oscillates rapidly even within a unit cell, very large $N$, around 50 times the number of unit cell $L/a_0$, is needed to achieve convergence of energy eigenvalues near valence or conduction band edge. \begin{figure}[t] \includegraphics[width=0.45\textwidth]{fig5} \caption{\label{fig:fig5} Valence and conduction band edge plotted as a function of position in inhomogeneously strained crystal. The local band edges are calculated from homogeneously strained crystal with the same strain magnitude at position $x$.} \end{figure} Fig.~\ref{fig:fig4}a shows the direct-diagonalization obtained energy eigenvalues near the band edges. A $5000\times 5000$ Hamiltonian matrix is involved in the numerical calculation. For comparison, the energy eigenvalues of unstrained crystal are shown together in the figure. The most distinct feature for the energy spectrum of inhomogeneously strained crystal is the appearance of bound states near the conduction and valence band edges. These bound states, whose wavefunctions are shown in Fig.~\ref{fig:fig4}b, can be understood by plotting the local valence and conduction band edges as a function of position in the strained crystal, which is shown in Fig.~\ref{fig:fig5}. The alignment of band edges is reminiscent of semiconductor quantum well, except that in our case, the spatial variation of band-edge is smooth and extended, while in semiconductor quantum well, band edge usually jumps abruptly at the interface between the barrier and well region of quantum well. Hence, the strain-confined bound states in inhomogeneously strained crystal bear resemblance to bound states in quantum well. We want to emphasize that, the band edge alignment in our 1D inhomogeneously strained crystal is not unique to this model. Strain-induced band edge shift in semiconductor is a well-known phenomenon.\cite{Bardeen50} In fact, the band-edge alignment in our 1D model is similar to those calculated by Feng \textit{et al} for inhomogeneously strained MoS$_2$ monolayer. \cite{Feng12} We can therefore conclude that the existence of electron or hole bound states is a general feature in an inhomogeneously strained crystal. \begin{figure}[t!] \includegraphics[width=0.4\textwidth]{fig6} \caption{\label{fig:fig6} Relative difference of energy eigenvalues obtained by direct diagonalization and envelope function method. The energy eigenvalues from direct diagonalization of a 5000 by 5000 Hamiltonian matrix are served as reference to calculate the relative difference. In (a), zone-center Bloch functions of the lowest five bands are used to carry out envelope function expansion. The envelope functions are represented numerically using one mesh grid every unit cell. This leads to the diagonalization of an approximately 500 by 500 matrix. In (b), only valence and conduction bands zone-center Bloch functions are involved in envelope function expansion. The envelope functions are represented using one mesh grid every four unit cells. The resulting matrix for diagonalization is of order 50 by 50.} \end{figure} We have also calculated the energy eigenvalues using our envelope function method. As shown in Fig.~\ref{fig:fig6}a, very high accuracy of eigenvalues is achieved for the whole valence and conduction bands using only one mesh grid per unit cell representation of the envelope functions. The lowest five bands are included in the summation over bands in the envelope function expansion (Eq.~\ref{eq:1D_envelope_expansion}). Together, the envelope function method involves the diagonalization of an approximately $500$ by $500$ matrix, which is an order of magnitude smaller than direct diagonalization. As zone-center Bloch functions are used to carry out envelope function expansion, naturally the error for energy eigenvalues near the band edge is smaller, same as in conventional envelope function method. Furthermore, if one is only concerned with energy levels near the band edge, which in most practical application is true, the expense of envelope function method can be reduced by another order of magnitude by including only the most relevant bands, and using coarser grids for numerical representation of the envelope functions. In Fig.~\ref{fig:fig6}b, we show that more than 1/4 of energy levels in valence and conduction bands can be calculated with very high accuracy by including only the valence and conduction bands in wavefunction expansion, and using one mesh grid every four unit cells to represent the envelope functions. In this case, one ends up diagonalizing a $50$ by $50$ matrix, which is two order of magnitude smaller than direct diagonalization. Indeed, for this 1D model, our envelop function method is much faster than the direct diagonalization method. \begin{figure}[t!] % \includegraphics[width=0.475\textwidth]{fig7} % \caption{\label{fig:fig7} Amplitude square plot of envelope functions $F_n(x)$ for states near valence and conduction band edges. The electronic states plotted are VBM, VBM$+1$, CBM and CBM$-1$. For these band edge states, only valence band ($n=2$) and conduction band ($n=3$) have significant envelope function amplitudes.} % \end{figure} % The success of the envelope function method is because the envelope functions $F_n(x)$ are indeed slowly varying as we conjectured. Fig.~\ref{fig:fig7} shows the amplitude square plot of envelope functions for a few electron and hole bound states. For the electron bound states, the envelope function of conduction band is predominant, while the valence band envelope function also contributes. The opposite is true for the hole bound states. Other remote bands have negligible contribution and are therefore not plotted. Comparing with the full wavefunctions calculated from direct diagonalization in Fig.~\ref{fig:fig4}b, one can notice that the envelope functions are indeed slowly-varying functions modulating the amplitude of fast-varying Bloch functions. \section{Toward Application to Three-Dimensional Real Materials \label{sec:3DIssues}} We have demonstrated in the previous section that our envelope function method can be successfully applied to a model 1D slowly-varying inhomogeneously strained semiconductor. A real semiconductor, however, is a three-dimensional (3D) object, and its crystal potential and strain response will be more complicated than the 1D model. Therefore, in this section we discuss some of the issues that may arise when applying our method to real 3D semiconductor crystals. \begin{figure*}[t!] \includegraphics[width=0.9\textwidth]{fig8} \caption{\label{fig:fig8} Flow chart to implement the envelope function method described in this article. The first step is the determination of a smooth displacement field $\bu(\bx)$ which can map the unstrained crystal (and the associated vacuum space, if any) to the strained crystal. The strain field $\varepsilon(\bx)$ can be calculated from the displacement field $\bu(\bx)$. The second and third steps are construction of strain-parametrized Bloch functions and energy eigenvalues at the reference crystal momentum (usually Brillouin zone center), through \textit{ab initio} or semi-empirical electronic structure calculations of a series of homogeneously strained crystal, using strain values taken from the inhomogeneously strained crystal. The last step is the solution of the coupled differential equation for the envelope functions as a generalized matrix eigenvector problem.} \end{figure*} The procedures to carry out our envelope function method in 3D are essentially the same as in 1D, which we summarize in the flow chart of Fig.~\ref{fig:fig8}. The first step in the flow chart is the determination of a smooth displacement field $\bu(\bx)$ which can map the unstrained crystal (and the associated vacuum space, if any) to the strained crystal. The corresponding strain field $\varepsilon(\bx)$ needs to be calculated as well. In 1D, displacement field $u(x)$ is a one-dimensional function and contains no rotational component. Thus $u(x)$ is related to the strain field $\varepsilon(x)$ via a simple integral relation $u(x) = \int^x\varepsilon(v)dv$. In 3D, the displacement field $\bu(\bx)$ is three-dimensional, and the strain field $\boldsymbol \varepsilon(\bx)$ is a tensor field with six independent components. Due to the possible existence of rotational components, the components of the strain field are related to the displacement field $\bu(\bx)$ (in the small deformation limit) as: %%%%% \begin{equation} \varepsilon_{ij} = \frac{1}{2}\left(\frac{\partial u_i}{\partial x_j} + \frac{\partial u_j}{\partial x_i} \right) \end{equation} %%%%% Hence, for a generic 3D inhomogeneously strained crystal, finding and representing the smooth displacement field $\bu(\bx)$ and strain field $\boldsymbol \varepsilon(\bx)$ becomes more difficult than 1D. Knowledge of solid mechanics will be helpful in this endeavor. The increased complexity of the displacement field and strain field in 3D also complicates the calculation and representation of the differential operators $\mathcal{P}^*$ and $\mathcal{P}_0^*$ defined in Eq.~\ref{eq:operator_P} and Eq.~\ref{eq:operator_P0}, which need to be determined in order to solve the envelope function equation, Eq.~\ref{eq:Coupled_Envelope_Function_Equation}. The second and third step in the flow chart are the construction of strain-parametrized Bloch functions and associated strain-parametrized energy eigenvalues at a reference crystal momentum, usually at the Brillouin zone center, through \textit{ab initio} or semi-empirical electronic structure calculation of a series of \textit{homogeneously} strained crystal. The strain values are coarsely taken from the inhomogeneous strain field $\varepsilon$, which in principle is sufficient as the strain field is slowly-varying in space. Nevertheless, in a generic 3D case this step will be challenging as the number of calculations for homogeneously strained crystal can become quite large if the strain field is complex, as there are six independent components of strain tensor in 3D. The construction of strain-parameterized Bloch functions, described in Sec.~\ref{subsec:parametrized_basis}, might also become non-trivial due to the complexity of electronic wavefunctions in 3D. Proper choice of expansion basis $\varphi_m(\bx)$ for the Bloch functions in Eq.~\ref{eq:Strained_Blochwave_Expansion} and Eq.~\ref{eq:Parametrized_Expansion_Basis} will be essential. The fourth step in the flow chart is the solution of coupled differential equation for the envelope functions, the Eq.~\ref{eq:Coupled_Envelope_Function_Equation}. This step, having been discussed in Sec.~\ref{subsec:locality}, should be straightforward once the differential operators $\mathcal{P}^*$ and $\mathcal{P}_0^*$, strain parameterized Bloch functions $\ubx$ and the associated strain parameterized energy eigenvalues $\epsilon_{n0}(\varepsilon(\bx))$ have all been determined in the previous steps. Nevertheless, the computational cost of solving the differential eigenvalue equation will become larger as the dimensionality of the problem increases, as more spatial or Fourier grids will be needed to represent the envelope functions $F_n(\bx)$, resulting in larger matrices for numerical diagonalization. In summary, the application of our envelope function method to a generic 3D problem will be feasible but challenging. We therefore believe that our method will most likely find applications in cases where the 3D problem is quasi-1D or 2D, namely when only one or very few components of the strain tensor is varying slowly in space. We also comment here a few issues related to the central approximation adopted in our method, the local approximation of crystal potential in strained crystal elaborated in Sec.~\ref{subsec:locality}. The approximation states that in a slowly-varying inhomogeneously strained semiconductor or insulator, the local crystal potential $V(\bx')$ can be well approximated by that of a homogeneously strained crystal with the same strain tensor $\varepsilon(\bx')$. This local approximation of strained crystal potential is likely to be a good approximation only for non-polar semiconductors such as silicon and germanium. For polar semiconductors such as gallium arsenide, strain could induce piezoelectric effect, which generates long-range electric field in the deformed crystal and significantly increases the error of this approximation. Furthermore, we note that for certain materials with more than one atoms within a unit cell, strain can induce internal relaxation of atoms relative to each other on top of the displacement described by strain tensor, an effect not included in our present method and must be carefully checked in realistic calculations. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%% Envelope Function Written in Empirical Form %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Envelope Function Equation for Empirical Applications \label{sec:empirical}} In this section, we will cast the envelope function equation (Eq.~\ref{eq:Coupled_Envelope_Function_Equation}) in a new form in which the strain-parametrized Bloch functions $\ubx$ will not appear explicitly. They will be replaced by a set of matrix elements involving their integrals. Doing so allows the method to be used empirically, where the matrix elements can be fitted to experimental data. The connection to traditional $\kp$ envelope function method will also become clearer. For convenience, we rewrite the relevant equations below %%%%% \begin{equation} \begin{split} & \sum_n\mathcal{P}^*\left[F_n(\bx) \ubx \right] - \sum_n F_n(\bx) \mathcal{P}_0^*\left[\ubx\right] \\ & =\sum_n F_n(\bx) \left[ E - \epsilon_{n0}(\varepsilon(\bx)) \right]\ubx, \end{split} \label{eq:Reproduced_Envelope_Function_Equation} \end{equation} %%%%% where %%%%% \begin{eqnarray} &&\mathcal{P}^*= -\frac{1}{2} a_{ij}(\bx)\frac{\partial ^2}{\partial x_i \partial x_j} -\frac{1}{2} b_i(\bx)\frac{\partial}{\partial x_i}, \\ &&\mathcal{P}_0^* = -\frac{1}{2}(I+\varepsilon(\bx))^{-1}_{im}(I+\varepsilon(\bx))^{-T}_{mj} \left.\frac{\partial ^2}{\partial x_i \partial x_j}\right|_{\varphi(\bx)}. \end{eqnarray} %%%%% $a_{ij}(\bx)$ and $b_{i}(\bx)$ are given by %%%%% \begin{eqnarray} && a_{ij}(\bx) = \left(I+\nabla \bu(\bx) \right)_{im}^{-1} \left( I+\nabla \bu(\bx) \right)_{mj}^{-T}, \\ && b_i(\bx) = \left( I+\nabla \bu(\bx) \right)_{nm}^{-1} \frac{\partial}{\partial x_n} \left(I+\nabla \bu(\bx) \right)_{mi}^{-T}. \end{eqnarray} %%%%% $\mathcal{P}^*\left[F_n(\bx) \ubx \right]$ can be expanded out as %%%%% \begin{equation} \begin{split} &\mathcal{P}^*[F_n(\bx) \ubx] = \\ & [\mathcal{P}^*F_n(\bx)] \ubx + F_n(\bx)[\mathcal{P}^* \ubx]\\ & - a_{ij}(\bx)\frac{\partial F_n(\bx)}{\partial x_i}\frac{\partial}{\partial x_j}\ubx. \end{split} \end{equation} %%%%% In above expansion, we have used the symmetry property of $a_{ij}(\bx)$, namely $a_{ij}(\bx) = a_{ji}(\bx)$. When strain variation $\varepsilon(\bx)$ is varying slowly at atomic scale, which is the premise of our envelope function method, the strain-parametrized basis functions $\ubx$ for different bands $n$ are approximately independent and orthogonal: %%%%% \begin{equation} \frac{1}{V} \int d \bx \J(\bx) \left[\ubx \right]^{\dagger} \ubxm \approx \delta_{mn}. \end{equation} %%%%% The integration is over the whole crystal, whose volume is $V$. The Jacobian of deformation map $J(\bx) = \det(I + \nabla \bu)$ takes into account the change of volume elements during coordinate transformation. We also note that, $J(\bx)$ can be absorbed into the basis functions by re-defining $\ubx$ as $J(\bx)^{1/2}\ubx$, and the whole formalism of our envelope function method will not change. This can be sometimes be more convenient for constructing strain-parametrized basis set. Using the above orthonormal relation, we can express $\mathcal{P}^* \left[ \ubx \right]$, $\mathcal{P}_0^*\left[\ubx \right]$, and $\frac{\partial}{\partial x_i} \ubx$ in terms of $\ubx$ as %%%%% \begin{eqnarray} &&\mathcal{P}^* \left[\ubx \right]= \sum_{n'} P_{nn'}\ubxp, \nonumber \\ &&\mathcal{P}_0^* \left[\ubx \right]= \sum_{n'} P_{nn'}^0\ubxp, \nonumber \\ &&\frac{\partial}{\partial x_i} \ubx = \sum_{n'} Q_{nn'}^i \ubxp, \nonumber \end{eqnarray} %%%%% where $P_{nn'}$, $P_{nn'}^0$ and $Q_{nn'}^i$ are matrix elements given by %%%%% % \begin{widetext} % \begin{eqnarray} && P_{nn'} = \frac{1}{V} \int d \bx \J(\bx) \left\{\mathcal{P}^*\left[\ubx \right]\right\}\left[\ubxp \right]^{\dagger}, \nonumber \\ && P_{nn'}^0 = \frac{1}{V} \int d \bx \J(\bx) \left\{\mathcal{P}_0^* \left[\ubx \right]\right \} \left[\ubxp \right]^{\dagger}, \nonumber \\ && Q_{nn'}^i = \frac{1}{V} \int d \bx \J(\bx) \left\{ \frac{\partial}{\partial x_i} \ubx \right \} \left[\ubxp \right]^{\dagger}. \nonumber \end{eqnarray} %%%%% Eq.~\ref{eq:Reproduced_Envelope_Function_Equation} can now be written in terms of $\ubx$ as %%%%% \begin{widetext} \begin{equation} \sum_n \left\{\mathcal{P}^*F_n - \sum_{n'} a_{ij}(\bx) Q_{n'n}^i \frac{\partial F_{n'}}{\partial x_j} + \sum_{n'} (P_{n'n} - P_{n'n}^0) F_{n'} \right\}u_{n0}^* = \sum_n F_n(\bx) \left[E - \epsilon_{n0}(\varepsilon(\bx)) \right]u_{n0}^*. \end{equation} %%%%% Equating coefficients of $u_{n0}^*$ on both side,\cite{Burt92} we arrives at a new form of envelope function equation %%%%% \begin{equation} -\frac{1}{2} a_{ij}(\bx) \frac{\partial^2 F_n}{\partial x_i \partial x_j} - \frac{1}{2} b_i(\bx) \frac{\partial F_n}{\partial x_i} - \sum_{n'} a_{ij}(\bx) Q_{n'n}^i \frac{\partial F_{n'}}{\partial x_j} + \sum_{n'} (P_{n'n} - P_{n'n}^0) F_{n'} + \epsilon_{n0}(\varepsilon(\bx)) F_n = E\,F_n\end{equation} \end{widetext} %%%%% In the equation, $a_{ij}(\bx)$ and $b_i(\bx)$ are related to deformation mapping and can be calculated once the displacement field $\bu(\bx)$ is known. $Q_{n'n}^i$ and $(P_{n'n} - P_{n'n}^0)$ can be calculated either by constructing the strain-parametrized basis set or fit empirically to experimental data. As a sanity check, when a crystal is undeformed, namely $\bu(\bx) = 0$, we have $a_{ij}(\bx) = \delta_{ij}$, $b_i(\bx) = 0$, $J(\bx) = 1$, $P_{nn'} = P_{nn'}^0$, $\epsilon_{n0}(\varepsilon(\bx)) = \epsilon_{n0}$, and the envelope function equation will become %%%%% \begin{equation} -\frac{1}{2} \nabla^2F_n - \sum_{n'} q_{n'n}^i \frac{\partial F_{n'}}{\partial x_i} + \epsilon_{n0} F_{n} = E\, F_n\label{eq:EnvelopeFunction_Bulk_Crystal} \end{equation} %%%%% with $q_{n'n}^{i}$ being %%%%% \begin{equation} q_{n'n}^i = \frac{1}{V} \int d \bx \left[u_{n0}(\bx) \right]^{\dagger} \frac{\partial}{\partial x_i} u_{n'0}(\bx) \end{equation} %%%%% Eq.~\ref{eq:EnvelopeFunction_Bulk_Crystal} recovers the envelope function equation for bulk crystals.\cite{Burt92} \section{Summary and Conclusion} To summarize, we have developed a new envelope function formalism for electrons in slowly-varying inhomogeneously strained crystals. The method expands the electronic wavefunctions in a smoothly deformed crystal as the product of slowly varying envelope functions and strain-parametrized Bloch functions. Assuming, with justifications, that the local crystal potential in a smoothly deformed crystal can be well approximated by the potential of a homogeneously deformed crystal with the same strain value, the unknown crystal potential in Schr\"{o}dinger equation can be replaced by the a small set of strain-parametrized Bloch functions and the associated strain-parametrized energy eigenvalues at a chosen crystal momentum. Both the strain-parametrized Bloch functions and strain-parametrized energy eigenvalues can be constructed from \textit{ab initio} or semi-empirical electronic structure calculation of homogeneously strained crystals at unit-cell level. The Schr\"{o}dinger equation can then be turned into eigenvalue differential equations for the envelope functions. Due to the slowly-varying nature of the envelope functions, coarse spatial or fourier grids can be used to represent the envelope functions, therefore enabling the method to deal with relatively large systems. Compared to the traditional multi-band $\kp$ envelope function method, our envelope function method has the advantage of keeping unit-cell level microstructure information since the local electronic structure information is obtained from \textit{ab initio} or EPM calculations. Compared to the conventional EPM method, our method uses envelope function formalism to solve the global electronic structure, therefore has the potential to reduce the computational cost. The method can also be used empirically by fitting the parameters in our derived envelope function equations to experimental data. Our method thus provides a new route to calculate the electronic structure of slowly-varying inhomogeneously strained crystals. \section{Acknowledgements} We acknowledge financial support by NSF DMR-1120901. Computational time on the Extreme Science and Engineering Discovery Environment (XSEDE) under the grant number TG-DMR130038 is gratefully acknowledged. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \bibliography{reference} \end{document} }

196 — 1410.7639

\caption{Views of the GC region centred on the Sgr~B region (represented by the black box) and smoothed to a common resolution of $0.2^\circ$ with the beam shown in the lower left-hand corner. % (a) \fermi count map obtained by admitting only \grays above 6~GeV, using a linear transfer function and the intensity runs from 0 to 15 counts/bin. % (b) Planck dust opacity map obtained from the \planck collaboration \citep{planck}. % This image has had a logarithmic transfer function applied to it to enhance the diffuse emission, and its scale runs from 0 to $4.5\times10^{-8}$.}

\caption{Images of residual \grays above 1~GeV obtained using (a) the standard Fermi diffuse background (gll\_iem\_v05\_rev1.fit and iso\_source\_v05\_rev1.txt) and (b) our dust and IC model templates, as discussed in the text.}

\caption{Spectral energy distribution of \gray from the Sgr B complex. % The grey bands represent the error of the \fermi measurement and consist of both systematic and statistic errors. % The stars (with associated errors) represent the diffuse \gray flux discovered by H.E.S.S. \citep{hessgc}, and the dot-dashed line is the power law obtained using the photon index and normalisation obtained from the original H.E.S.S. paper reporting its discovery. % The dotted line represents the predicted \gray spectrum obtained by assuming that the CR spectrum in the Sgr~B region is the same as the local CR spectrum measured by PAMELA. }

\caption{Spectrum of different components of \gray emission from the Sgr~B region. % The (red) crosses represent the IC component, whilst the (blue) stars represent the isotropic component, both with associated errors shown. % The (black) plus signs are the total observed \gray component from the region. }

197 — 1410.8410

\caption{ \label{figs1} Data for $\beta = 2.2$ and $V = 48^4$ ({\bf \textcolor{red}{$+$}}) matched \cite{Leinweber:1998uu,Cucchieri:2003di} with data for $\beta = 2.34940204$ and $V = 72^4$ ({\bf \textcolor{green}{$\times$}}), fitted using Eq.\(\protect\ref{eq:fit}) with $t = 3.2(0.3) (GeV^2)$, $u = 3.6(0.4) (GeV)$, $s = 46(13) (GeV^2)$ and $c = 114(13)$. }

\caption{ \label{figs2} Data for $\beta = 2.34940204$ and $V = 72^4$ ({\bf \textcolor{green}{$\times$}}) matched \cite{Leinweber:1998uu,Cucchieri:2003di} with data $\beta = 2.43668228$ and $V = 96^4$ ({\bf \textcolor{blue}{$*$}}), fitted using Eq.\(\protect\ref{eq:fit}) with $t = 3.0(0.2) (GeV^2)$, $u = 3.9(0.3) (GeV)$, $s = 58.0(9.8) (GeV^2)$ and $c = 247(16)$. }

198 — 1410.8417

\caption{(Color online) Collective polariton condensate spin precession in a quasi 1D ridge. (a) Polariton PL distribution in real space under non-resonant (1.612 eV), circular-polarized ($\sigma^+$) excitation at the center of the ridge. The pump power is $3.75 \times P_{th}$. Dashed white arrows sketch the direction of the polariton flow along its propagation. (b) Corresponding circular degree of polarization distribution ($s_z$). Vertical blue and red arrows highlight the spin precession, oscillating from negative ($\sigma^-$) to positive ($\sigma^+$) values, respectively. (c) FFT intensity of the region enclosed by a dashed, white rectangle in panel (a). The dot-dashed, white rectangle marks the region of relevant frequencies arising from interferences between propagating and backscattered polaritons in real space. \redc{The PL and the FFT map are coded in a logarithmic, false color scale, while a linear one is used for $s_z$.} (d) PL ($s_z$) versus $x$ in the central region of the ridge $y=0$, plotted with a thick-gray (thin-purple) line.}

\caption{(Color online) Simulations of the collective polariton condensate spin precession in a quasi 1D ridge. (a) PL rendering polariton distribution in real space under non-resonant, circularly polarized excitation at the center of the ridge\redc{, with a pump power $P=3.75\times P_{th}$. The dashed, white box marks the spatial region that is Fourier transformed.} (b) Corresponding simulation on the circular degree of polarization distribution ($s_z$). \redc{(c) FFT intensity of the simulated polariton PL distribution in the framed area in panel (a); remarkable Fourier frequencies arise at $\sim3$ $\mu$m$^{-1}$ from the counter propagating polariton populations, see the delimited region by a dot-dashed, white box. The PL and the FFT map are coded in a logarithmic, false color scale, while a linear one is used for $s_z$.}}

\caption{(Color online) Simulation on the Stokes parameters of the polariton PL as function of energy and spatial position ($x$): (a) $s_x$, (b) $s_y$, and (c) $s_z$, respectively, under non-resonant, circular-polarized excitation\redc{, with a pump power $P=3.75\times P_{th}$. (d) $s_z$ under a higher pump power excitation $4.75\times P_{th}$.} The degree of polarization is coded in a linear, false color scale.}

199 — 1410.8541

\caption{All Possible Redundancy Allocation Candidates of $\left[ n = 1023, k=923, l, r \right]$ PBCH Codes} \label{tab:PLBC} \centering {\small \begin{tabular}{|c|c|c|c|} \hline Code & {$l$} & {$r$} & Notes \\ \hline \hline 0 & 0 & 100 & Only correcting random errors \\ \hline 1 & 10 & 90 &\\ \hline 2 & 20 & 80 &\\ \hline 3 & 30 & 70 &\\ \hline 4 & 40 & 60 &\\ \hline 5 & 50 & 50 &\\ \hline 6 & 60 & 40 &\\ \hline 7 & 70 & 30 &\\ \hline 8 & 80 & 20 &\\ \hline 9 & 90 & 10 &\\ \hline 10& 100 & 0 & Only masking defects\\ \hline \end{tabular}} \vspace{-5mm} \end{table} In this section, the simulation results are presented. The simulation parameters are summarized in Table.~\ref{tab:parameters}. The initial threshold voltage distribution (after erasing a flash memory block) is assumed to be the Gaussian distribution $\mathcal{N} \left(-4, 1^2\right)$. The ISPP was implemented with the parameters of the verify level for $S_1$, i.e., $\nu_{S_1} = 1$ and the incremental step voltage $\Delta V_\textrm{pp} = 1$. %Note that the variance of initial threshold voltage distribution and the incremental step voltage work for $Z_\textrm{write}$ of \eqref{eq:fmc1}, which precedes the ICI. The noise due to fast detrapping $Z_\textrm{fast}$ is assumed to be the Gaussian distribution $\mathcal{N} \left(-0.2, \sigma_{Z_\textrm{fast}}^2 \right)$ by taking into account experimental results in~\cite{Chen2010}. The random noise due to other read noise sources $Z_\textrm{random}$ is assumed to the $\mathcal{N} \left(0, \sigma_{Z_\textrm{random}}^2 \right)$. For the additive encoding, we use $\left[ n = 1023, k=923, l, r \right]$ partitioned Bose, Chaudhuri, Hocquenghem (PBCH) codes where $n$, $k$, $l$, and $r$ denote the codeword size, the information size, the redundancy size for masking defects, and the redundancy size for correcting random errors, respectively. The PBCH code is a special class of partitioned linear block codes, which can be designed by a similar method of standard BCH codes \cite{Heegard1983plbc}. For the given $n=1023$ and $k=923$ (i.e., the total redundancy size is 100), all possible redundancy allocation candidates of PBCH codes are presented in Table~\ref{tab:PLBC}~\cite{Kim2013redundancy}. Note that $l$ and $r$ are multiples of 10 that is the degree of Galois field. For the efficient computation complexity, two-step encoding scheme of \cite{Kim2013coding, Kim2013redundancy} has been used for encoding of PBCH codes. \begin{figure}[!t] \centering \includegraphics[width=3in]{fig_distribution.eps} \caption{Comparison of threshold voltage distributions ($\sigma_{{Z_\textrm{fast}}} = 0.4$, $\sigma_{{Z_\textrm{random}}} = 0.2$, $\zeta = 0.4$).} \label{fig:distribution} % \vspace{-5mm} \end{figure} \begin{figure}[!t] \centering \includegraphics[width=3in]{fig_wer_id.eps} \caption{$P(\text{decoding failure})$ for different identify levels $\zeta$ ($\sigma_{{Z_\textrm{fast}}} = 0.4$, $\sigma_{{Z_\textrm{random}}} = 0.2$).} \label{fig:plot_wer_id} \vspace{-5mm} \end{figure} \begin{figure}[t] \centering \subfloat[Comparison of $P(\text{decoding failure})$ for different $\sigma_{Z_\textrm{fast}}$]{\includegraphics[width=3in]{fig_wer_fast_a.eps} \label{fig:plot_wer_fast_a}} \hfil \subfloat[Comparison of $P(\text{decoding failure})$ for $(l=0, r=100)$, (i.e., not using any side information) and that for the optimal redundancy allocation $\left(l^*, r^*\right)$]{\includegraphics[width=3in]{fig_wer_fast_b.eps} \label{fig:plot_wer_fast_b}} \caption{$P(\text{decoding failure})$ for different $\sigma_{Z_\textrm{fast}}$ ($\sigma_{{Z_\textrm{random}}} = 0.2$, $\zeta = 0.2$).} \label{fig:plot_wer_fast} \vspace{-5mm} \end{figure} \begin{figure}[t] \centering \subfloat[Comparison of $P(\text{decoding failure})$ for different $\sigma_{Z_\textrm{random}}$]{\includegraphics[width=3in]{fig_wer_random_a.eps} \label{fig:plot_wer_random_a}} \hfil \subfloat[Comparison of $P(\text{decoding failure})$ for $(l=0, r=100)$, (i.e., not using any side information) and that for the optimal redundancy allocation $\left(l^*, r^*\right)$]{\includegraphics[width=3in]{fig_wer_random_b.eps} \label{fig:plot_wer_random_b}} \caption{$P(\text{decoding failure})$ for different $\sigma_{Z_\textrm{random}}$ ($\sigma_{{Z_\textrm{fast}}} = 0.4$, $\zeta = 0.2$).} \label{fig:plot_wer_random} \vspace{-5mm} \end{figure} Fig.~\ref{fig:distribution} shows that controlling ICI by additive encoding can compensate the effect of fast detrapping. After compensating the fast detrapping by the intentional ICI, the threshold voltage distribution improves. Fig.~\ref{fig:plot_wer_id} shows that the probability of decoding failure $P(\textrm{decoding failure})$ is improved by the proposed scheme. If the redundancy for masking $l$ is zero, it means that the side information of fast detrapping is ignored. Otherwise, the encoder uses the side information of fast detrapping to improve $P(\textrm{decoding failure})$. Note that $P(\text{decoding failure})$ depends on the identify level $\zeta$ since it controls the number of identified cells, which is equivalent to the number of cells identified as defects in the upper WL. Also, the optimal redundancy allocation $\left(l^*, r^*\right)$ to minimize $P(\text{decoding failure})$ depends on the identify level $\zeta$. For larger $\zeta$, the number of cells identified as defects increases, which requires more redundancy for masking defects during additive encoding. For the given parameters in Fig.~\ref{fig:plot_wer_id}, the identify level of $\zeta = 0.2$ minimizes $P(\textrm{decoding failure})$. Fig.~\ref{fig:plot_wer_fast} shows the improvement of $P(\textrm{decoding failure})$ by the proposed scheme for different levels of fast detrapping $Z_{\textrm{fast}}$. Note that $\sigma_{{Z_\textrm{fast}}}$ changes from 0.4 to 0.5 for the given $\sigma_{{Z_\textrm{random}}} = 0.2$. From Fig.~\ref{fig:plot_wer_fast}~\subref{fig:plot_wer_fast_a}, we can obtain the optimal redundancy allocation $\left(l^*, r^*\right)$ minimizing $P(\textrm{decoding failure})$. It is worth mentioning that the optimal redundancy for masking $l^*$ increases for larger $\sigma_{{Z_\textrm{fast}}}$. This is because we should allot more redundancy for masking defects (i.e., compensating the fast detrapping) as $\sigma_{{Z_\textrm{fast}}}$ increases. Fig.~\ref{fig:plot_wer_fast}~\subref{fig:plot_wer_fast_b} compares $P(\textrm{decoding failure})$ for $(l, r) = (0, 100)$, (i.e., coding without the side information of fast detrapping) and that for $\left(l^*, r^*\right)$ (i.e., coding with the side information of fast detrapping). By using the side information of fast detrapping, we can significantly improve $P(\textrm{decoding failure})$ for different $\sigma_{{Z_\textrm{fast}}}$. Fig.~\ref{fig:plot_wer_random} shows the improvement of $P(\textrm{decoding failure})$ by the proposed coding scheme for different random noise $Z_{\textrm{random}}$. Note that $\sigma_{{Z_\textrm{random}}}$ changes from 0.2 to 0.3 for the given $\sigma_{{Z_\textrm{fast}}} = 0.4$. From Fig.~\ref{fig:plot_wer_random}~\subref{fig:plot_wer_random_a}, we can obtain the optimal redundancy allocation $\left(l^*, r^*\right)$ minimizing $P(\textrm{decoding failure})$. Fig.~\ref{fig:plot_wer_random}~\subref{fig:plot_wer_random_b} compares $P(\textrm{decoding failure})$ for $(l, r) = (0, 100)$, (i.e., coding without the side information of fast detrapping) and that for $\left(l^*, r^*\right)$ (i.e., coding with the side information of fast detrapping). It is worth mentioning that the improvement of $P(\textrm{decoding failure})$ becomes significant as the fast detrapping $Z_{\textrm{fast}}$ dominates the random noise $Z_{\textrm{random}}$. \section{Conclusion}\label{sec:conclusion} We proposed a scheme to compensate the effect of fast detrapping by intentional ICI. The main idea comes from the observation that ICI increases the threshold voltage of a cell whereas fast detrapping decreases the threshold voltage of corresponding cell. Additive encoding can control the intentional ICI by using the side information of fast detrapping. Although this paper focused on SLC flash memory, the proposed scheme can be extended to MLC flash memories. %% Appendix: %% If needed a single appendix is created by %\appendix %% If several appendices are needed, then the command %\appendices %% in combination with further \section-commands can be used. %\appendix %%% Use \section* for acknowledgement %\section*{Acknowledgment} % %The authors would like to thank various sponsors for supporting %their research. %% References: %% We recommend the usage of BibTeX: %% \IEEEtriggeratref{5} \bibliographystyle{IEEEtran} \bibliography{IEEEabrv,fast_detrap} \end{document} ]\end{figure}}

200 — 1410.8641

\caption{\blue{Same as \Fig{figalmc} but in this model, the large-scale flow has three cells radially stacked in the solar convection zone and the differential rotation includes near surface shear layer.} }

201 — 1410.8859

\caption{GP fit for the change in drag from the baseline (no-control) case $\Delta D/D_{baseline}$ versus wave speed $c$. Negative values of $c$ correspond to upstream traveling waves. The region bounded by one standard deviation $\sigma$ of the GP is shaded grey. (\dashed) GP mean,(\solidcircle) evaluations, (\starsol) snapped evaluations. \label{f:channel_GP} }

\caption{(a) Computational mesh around the baseline geometry. (b) Baseline (\dashed) and optimal(\solid) trailing edges. (\solidcircle) indicates fixed points and (\starsol) control points used in the design. \label{f:trailing_edge_geom} }

202 — 1411.0539

\caption{a) Mucosa dataset, window $[0, 1] \times [0.05, 0.75]$ type 1 =$\circ$, type 2 = {\color{blue}$\times$}. b) Posterior pointwise 95\%-envelopes for the trends without intercept, as a function of $y$-coordinate.}

\caption{a) Interaction function estimation using a step-function with a smoothing prior. True interaction used in simulation ({\color{blue}blue}, thick line), 1000 posterior simulations of the step function, LOESS-smoothed (light gray lines), and the posterior means of the step function weights ({\color{darkred}red $\times$}). b) Interaction function estimated using 50 square exponential basis functions.}

203 — 1411.0581

\caption{(Color online) Normalized differential conductance of a point contact in the superconducting state; URu$_2$Si$_2$ - Au(W). Panel A shows the rough data measured from {\color{red} 1.7} to 0.3 K. The gap value versus temperature is displayed in B. The evolution does not follow the BCS theory. In order to see the influence of the Fano resonance we subtracted from these data the curve measured at 1.7 K, the resultant curves are shown in panel C, there we plotted the differential conductance normalized at 1.5 K. In D is plotted the superconducting energy gap as a function of normalized temperature, this data follow quite well the BCS theory. This strong and different evolution with temperature of the gap is due to the Fano resonance.}

204 — 1411.0606

\caption[Clustering of coffee data samples]{Projection of coffee data samples marked according to the clustering obtained from the variables selected using the forward/backward greedy search. The symbol \textcolor{green3}{\sf R} indicates Robusta coffees, \textcolor{dodgerblue2}{\sf A} and \textcolor{red3}{\sf a} the sub-varieties of Arabica coffees.}

205 — 1411.0957

\caption{(a) A time-series of the angular position of the magnetic-dipole axis of the sphere measured as a function of time as the sphere rotates back to the zero-field orientation under the sole influence of gravity. A least-squares fit of the data to $\theta = A \exp(-t/T)+B$ yielded $A=28.21^\circ$, $B=3.09^\circ$ and $T_{0}=4.67$ s. (b) The phase delay, $\varphi$, between the applied magnetic field and the resulting motion of the sphere given by $\varphi=\arctan(2\pi T_0 f)$, depicted by the solid, black line, and measured experimentally (\textcolor{blue}{o}) as a function of the frequency of applied field, for values of $Ma$ and $\hat{\epsilon}$ such that $\theta \approx 16^{\circ}$. (c) Experimental time-series of the angular displacement of the driven, magnetic sphere over four periods of oscillation, for $\hat{\epsilon}=0.22$ and $Ma = 4.85$, $2.43$, $1.21$, $0.48$ and $0.20$ (top-to-bottom). The response of the sphere deviates increasingly from the form of the sinusoidal drive with decreasing $Ma$.}

\caption{The total angular displacement of the sphere measured as a function of the dimensionless parameter $Ma$ for various $\hat{\epsilon}$. Circles (\textcolor{blue}{o}) represent the experimental data and crosses (\textcolor{red}{x}) represent the numerical solutions evaluated at the corresponding parameters. Error bars on the experimental data points have not been included to aid visual clarity. The analytic solution for the case when the gravitational torque acting on the sphere is zero, $\hat{\epsilon}=0$, is represented by the solid, black line. Numerical results obtained for $\hat{\epsilon}=0.44$, $\hat{\epsilon}=0.67$, $\hat{\epsilon}=1.33$ and $\hat{\epsilon}=3.34$ are represented as indicated in the legend.}

\caption{The instantaneous fluid velocity measured as a function of radial distance. The fluid velocity $v$ has been normalised by the surface velocity of the sphere $v_S$, and the radial distance $r$ has been normalised by the radius of the sphere $a$. The black curve is the analytic solution for the fluid velocity due to a sphere rotating with constant angular velocity in an infinite fluid \cite{Jeffery1915}. The blue data points (\textcolor{blue}{o}) represent the experimental velocity, the standard deviation of which is given by the error bars. Measurements were made for $Ma = 2.01$ and $\hat{\epsilon}=0.22$ for which the sphere performed torsional oscillations of amplitude $\theta = 28^{\circ}$ that approximated a sinusoidal function to within $1\%$.}

206 — 1411.1147

\caption{Left: Examples of structured observations (in black), hidden structures (in {\color[rgb]{0.5,0.5,0.5} gray}), and side information (\underline{underlined}). Right: Model variables for POS induction (\S\ref{sec:pos}) and word alignment (\S\ref{sec:wa}). A parallel corpus consists of pairs of sentences (``source'' and ``target''). \label{fig:examples} \label{table:spec}}

207 — 1411.1223

\caption{Two significant components of the structural sensitivity tensor, which quantify the effect of feedback mechanisms, placed at $x$, on the linear growth rate (left frames) and angular frequency (right frames). Solutions with no mean-flow temperature jump (solid lines) and with temperature jump (dash-dotted line). $c_1=0.01$, $c_2=0.001$, $\tau=0.01$, $\beta=0.433$, $x_f=0.25$, \textcolor{\ccolor}{10 Galerkin modes are used for numerical discretization}.}

208 — 1411.1445

\caption[]{A range of EBL intensities from the UV to sub-millimeter wavelengths is shown (from Dwek \& Krennrich 2013 and references therein). The blue squares indicate lower limits from galaxy counts. The red circles show absolute measurements. The shaded area indicates the EBL intensity allowed by a wide range of UV to sub-millimeter observations.}

\caption[]{The difference between $\rm \Gamma_{GeV}$, the spectral index at GeV (\fermilat) energies, and $\rm \Gamma_{TeV}$, the energy spectral index in the TeV regime (H.E.S.S, MAGIC, VERITAS) is shown as a function of their redshift. Red squares (radio galaxies), red stars (starburst galaxies), empty circles (HBLs, high-frequency peaked BL Lacs), blue downward triangles (intermediate-frequency peaked BL Lacs), filled circles (LBLs, low-frequency peaked BL Lacs), red upward triangles (FSRQs, flat spectrum radio quasars) indicate the different types of \gray\sources. The figure has been adapted from Dwek\& Krennrich 2013, however, a recent data point of PKS~1424+240 has been added (blue upward triangle), with the caveat that the redshift of the source is a lower limit.}

209 — 1411.1651

\caption{Ensembles used in these calculations. \red{Red} indicates physical quark mass ensembles.}{ \begin{tabular}{c c c c c c c c c}\hline\hline \multirow{2}{*}{$a$ (fm)} & \multirow{2}{*}{$m_l/m_s$} & \multirow{2}{*}{Volume} & \multicolumn{2}{c}{$N_{conf}\times N_{t_{source}}$}& \multirow{2}{*}{$am_s^{sea}$} & \multirow{2}{*}{$am_s^{val}$} & \multirow{2}{*}{$am_c^{sea}$} & \multirow{2}{*}{$am_c^{val}$} \\ & & & ($K\to\pi$) & ($D\to K/\pi$) & & & & \\ \hline 0.15 &\red{0.035} & $32^3\times 48$ & $1000\times 4$ & & \red{0.0647} & \red{0.0691} & 0.831 & 0.8531 \\ \hline 0.12 & 0.2 & $24^3\times 64$ & $1053\times 8$ & $1050 \times 8$ &0.0590 & 0.0535 & 0.635 & 0.6363 \\ & 0.1 & $32^3\times 64$ & $993 \times 4$ & $993 \times 4$ & 0.0507 & 0.053 & 0.628 & 0.650 \\ & 0.1 & $40^3\times 64$ & $391 \times 4$ & &0.0507 & 0.053 & 0.628 & 0.650 \\ & \red{0.035} & $48^3\times 64$ & $945 \times 8$ & $943 \times 8 $ &\red{0.0507} & \red{0.0531} & 0.628 & 0.6269 \\ \hline 0.09 & 0.2 & $32^3\times 96$ & $755 \times 4$ & $773 \times 4 $& 0.037 & 0.038 & 0.45 & 0.44 \\ & 0.1 & $48^3\times 96$ & $853 \times 4$ & $851 \times 4$ & 0.0363 & 0.038 & 0.44 & 0.43 \\ & \red{0.035} & $64^3\times 96$ & $963 \times 8$ & $905 \times 8$ &\red{0.0363} & \red{0.0363} & 0.432 & 0.432 \\ \hline 0.06 & 0.2 & $48^3\times 144$& $362 \times 4$ & &0.024 & 0.024 & 0.286 & 0.286 \\ & \red{0.035} & $96^3\times 192$& $565 \times 6$ & $565 \times 6$ &\red{0.022} & \red{0.022} & 0.26 & 0.26 \\ \hline \hline \end{tabular}\label{ensembles}}

210 — 1411.1671

\caption{\label{fig:OnShellContrast}\textbf{Amplitude of the CBS revivals $\gamma_\mathrm{rev}(t)$ for several dephasing times $T$}. The data points \TriangleUp \textcolor{red}{\CircleSolid} \textcolor{blue}{\SquareSolid} \textcolor{OliveGreen}{\DiamondSolid} correspond to $T=$~0.7, 1.0, 1.3, and 1.6~ms respectively, the kick's strength $\Delta k$ being fixed (same as in Fig.~\ref{fig:CoherentPartEvolution}). The dotted vertical lines indicate times $2 T$ when the revivals are expected. Solid lines are Gaussian fits with zero offset. The observed revival times, determined by the fits, are respectively (in ms): 1.3 $\pm$ 0.08, 1.99 $\pm$ 0.08, 2.65$\pm$ 0.05 and 3.18$\pm$ 0.09. Uncertainties correspond to the 95\% confidence intervals.}

211 — 1411.2541

\caption{(color online) NCSMC n-\elem{Be}{8} phase shifts and eigenphase shifts for negative parity at $N_{\text{max}}=8$ computed with NCSM eigenvectors (\linemediumsolid) compared to the use of IT-NCSM vectors (\linemediumdashed). Remaining parameters are $\hbar\Omega=20\,\text{MeV}$, $\alpha=0.0625\,\text{fm}^4$, and $E_{\text{3max}}=14$. Same colors correspond to identical angular momenta. }

212 — 1411.3172

\caption[]{\label{fig::Corr_Sigma_B}\red{Pearson correlation matrix corresponding to $\bs\Sigma_{VB}$ for the 21 experiments (Cases 1-4 of LCT-007 and Cases 1-17 of LCT-039).}}

\caption[]{\label{fig::prior_post_bench}\red{Mean values of the prior and posterior $k_{\rm eff}$ distributions for the 21 benchmark experiments. The error bars show the standard deviations of the distributions.}}

\caption[]{\label{fig::post_comvergence}\red{Posterior $k_{\rm eff}\pm \sigma$ for the application case as a function of the number of benchmark experiments taken into account in the Bayesian updating procedure.}}

213 — 1411.3217

\caption{\label{fig:rvcurv} Top: The radial velocity curve measured from the Tb~\textsc{iii} $\lambda\,5505$~\AA\line. The solid curve is a least-squares fit using the average frequency of 1434~$\mu$Hz. Bottom: This is the same radial velocity curve fitted with the two photometric frequencies of $1411~\mu$Hz and $1452~\mu$Hz.}

\caption[]{Pulsation amplitudes and phases from radial velocity measurements for various ions in HD~213637. The phases were calculated for the frequency 1434~$\mu$Hz and a time zero point of MJD~53892.050. No phase is given for ions for which the amplitude is not significant. \red{The ions are listed in order of increasing pulsation amplitude.}}

214 — 1411.3272

\caption{In the real case~\eqref{eq:P_real}, one can hope to recover the signal $z \in \{\pm 1\}^n$ exactly from the pairwise sign comparisons $zz^{\!\top} + \sigma W$. This figure shows how frequently the semidefinite relaxation~\eqref{eq:SDP_real} returns the correct signal $z$ (or $-z$). For each pair $(n, \sigma)$, 100 realizations of the noise $W$ are generated independently and the dual certificate for the true signal~\eqref{eq:S_real} is verified (it is declared numerically positive semidefinite if its smallest eigenvalue exceeds $-10^{-14}n$). The frequency of success is coded by intensity (bright for 100\% success, dark for 0\% success). The results are in excellent agreement with the theoretical predictions~\eqref{eq:real_rate} (\textcolor{mycolor1}{red curve}).}

\caption{In the complex case~\eqref{eq:P}, \emph{exact} recovery of the phases $z$ from the pairwise relative phase measurements $zz^* + \sigma W$ is hopeless as soon as $\sigma>0$. Nevertheless, below the \textcolor{mycolor1}{middle line}, the perturbation is smaller than the signal (in operator norm, assuming $z$-discordance), and one expects to be able to recover $z$ reasonably well. And indeed, the maximum likelihood estimator (MLE) for $z$ is close to $z$, as per Lemmas~\ref{lemma:Delta_L2bound} and~\ref{lem:deltainfty}. Computing the MLE is hard in general, but solving the semidefinite relaxation~\eqref{eq:SDP} is tractable. When~\eqref{eq:SDP} has a rank-one solution, that solution coincides with the MLE. This figure shows, empirically, how frequently the \eqref{eq:SDP} admits a unique rank-one solution (same color code as Figure~\ref{fig:realrecovery}). For each pair $(n, \sigma)$, 100 realizations of the noise $W$ are generated independently and \eqref{eq:SDP} is solved using a complex version of the low-rank algorithm in~\cite{boumal2015staircase}. The~\eqref{eq:SDP} appears to be tight for remarkably large levels of noise. Theorem~\ref{Thm:rr_gaussian} (our main contribution) partly explains this phenomenon, by showing that the SDP is tight with high probability below the \textcolor{mycolor3}{bottom line}. We further note that, above the \textcolor{mycolor2}{top line}, no unbiased estimator for $z$ performs better than a random guess~\cite{BoumalManifoldSynch}. Note that, for $n \leq 3$, a complex version of~\cite[Thm.~2.1]{pataki1998rank} guarantees deterministic tightness of the relaxation, in accordance with observation here.}

215 — 1411.3312

\caption{Part of the $(3,4)$-nuclei forest for \notredame. In the entire forest, there are $2059$ nodes and $812$ leaves. $79$ of the leaves are clique, up to the size of $155$. There is a nice branching structure leading to a decent hierarchy.}

\caption{Density histograms for nuclei of three graphs. $x$-axis (binned) is the density and $y$-axis is the number of nuclei (at least 10 vertices) with that density. Number of nuclei with the density above 0.8 is significant: $139$ for \epinion, $355$ for \notredame, and $1874$ for \wiki. Also notice that, the mass of the histogram is shifted to right in \epinion{} and \wiki graphs.\label{fig:density}}

\caption{Overlap scatter plots for $(3,4)$-nuclei. Each axis shows the edge density of a participating nucleus in the pair-wise overlap. Larger density is shown on the $y$-axis. $(3,4)$-nuclei is able to get overlaps between very dense subgraphs, especially in \notredame{} and \wiki. In \wiki{} graph, there are $1424$ instances of pair-wise overlap between two nuclei, where each nucleus has the density of at least $0.8$. \label{fig:over-scatter}}

216 — 1411.3715

\caption{\label{fig:ranking}Mean values and confidence intervals of the accuracy of methods for \Acronym{ASC} evaluated on the \Acronym{DCASE} private dataset using stratified 5-fold cross-validation. The boxes enclose methods that cannot be judged to perform differently with a significance level of $95\%$. Please see Table \ref{tab:dcase} for the definition of the algorithms' acronyms. \textcolor{red}{MV is a majority vote classifier which assigns to an audio recording the label that is most commonly returned by the other methods. `H' indicates the median human accuracy, as obtained through the test described in Section \ref{sec:hp}, while `[31]' refers to the human accuracy obtained by Krijnders and Holt. Note that the confidence intervals displayed for the algorithmic results are not directly comparable to the variations in human performance, and hence only the median human performance is depicted. See Figure \ref{fig:hacc} for more details on the distribution of human accuracies.}}

\caption{\label{fig:sca}Distribution of algorithmic soundscapes classification accuracies. The solid line in the upper plot represents the average accuracy calculated from all the acoustic scenes. \textcolor{red}{The bottom plot depicts the histogram of mean accuracies resulting from the classification of all $100$ soundscapes, highlighting in the left tail that ten soundscapes correctly classified by at most only 10\% of the algorithms.}}

217 — 1411.4026

\caption{$\DC$ profiles of the semileptonic asymmetries from \cite{Botella:2014qya}; the {\color{blue} blue} line is the non $3\times 3$ unitary NP scenario -- eqs. \refeq{eq:NoUn:01} and \refeq{eq:NoUnMixq:01} --, the {\color{red} red dotted} line is the $3\times 3$ unitary NP scenario of \refeq{eq:33NP:Mix01} and the {\color{red} red dashed} line, the SM case. The D0 measurement is $\aslb=(-4.96\pm 1.69)\cdot 10^{-3}$ \cite{Abazov:2013uma}. \label{fig:NoUnNP:01}}

\caption{$\DC$ 68\%, 95\% and 99\% CL regions from \cite{Botella:2014qya}. {\color{blue} Blue} regions correspond to the non $3\times 3$ unitary NP scenario, {\color{red} red} regions correspond to the SM case.\label{fig:NoUnNP:02}}

218 — 1411.4102

\caption{\label{fig:singleparamcomp}AAAMS preserves more details and affects more perceptually salient segmentations, at similar clustering levels. We used a single parameter set,~{\tiny{}$\bigl\langle{\sigma_{base}^{r}}^{2},{\sigma_{base}^{s}}^{2},\,\epsilon_{r}^{2},\,\epsilon_{s}^{2}\bigr\rangle=\bigl\langle15,16,1,81\bigr\rangle$}{\tiny} with {\tiny{}$d_{B}=1$}{\tiny}, to show its adaptivity on varied images. JMS segments were kept around the same, with eye on preserving detail; it still smooths over at places. Its parameter values varied significantly from image to image - {\tiny{}$\,{\sigma^{r}}^{2}\in\left[49,\,81\right]\,,\,{\sigma^{s}}^{2}\in\left[100,\,289\right]$}{\tiny}. Minimum cluster size was {\tiny{}10}{\tiny}.} \end{minipage} \end{figure} \begin{figure}[h] \begin{minipage}[t!]{0.49\columnwidth} \vspace*{-.4mm} \tiny \noindent\begin{raggedright}~~~~~~~\emph{Image}~~~~~~~~~~~~~~~~~~~~~\emph{JMS}~~~~~~~~~~~~~~~~~~~~\emph{AAAMS}~~~~~~~~~~~~~~~~\emph{JMS~Labels}~~~~~~~~~\emph{AAAMS~Labels} \par\end{raggedright}\vspace*{0pt} \noindent\includegraphics[width=1\columnwidth]{aeroplane_compare}\\[-1pt] \noindent\raggedright{}\emph{\hspace*{3mm}Aeroplane~~~~~~~~~~~~~~~JMS - 14 Labels,~~~~~~~~~~~~AAAMS - 11 Labels} \noindent\includegraphics[width=1\columnwidth]{color_wheel_compare}\\[-1pt] \noindent\raggedright{}\emph{\hspace*{3mm}ColorWheel~~~~~~~~~~~~~~~JMS - 51 Labels,~~~~~~~~~~~~AAAMS - 48 Labels} \noindent\includegraphics[width=1\columnwidth]{wasp_compare}\\[-1pt] \noindent\raggedright{}\emph{\hspace*{3mm}Wasp~~~~~~~~~~~~~~~JMS - 150 Labels,~~~~~~~~~~~~AAAMS - 111 Labels} %\noindent\includegraphics[width=1\columnwidth]{compare_joined}\\[0pt] \caption*{\label{fig:parsimonycomp}\scriptsize{}Figure 4: More parsimonious segmentations were quite often not achievable with JMS - some varied examples are shown above (Images such as \emph{Lady} in \emph{Fig.3} are a typical case too). Both methods were configured for reduced label usage. Minimum cluster size was {\tiny{}10}{\tiny}. JMS, at its limit, is breaking boundaries and under segmenting. AAAMS with lesser labels, does not break boundaries, still maintains segment saliency.} \end{minipage} \hspace*{1mm} \begin{minipage}[t!]{0.48\linewidth} \noindent\includegraphics[width=1\columnwidth]{bear_clusters}\\[2pt] \noindent\includegraphics[width=1\columnwidth]{2D3Dsimulation}\\[-12pt] \vspace*{-1mm} \caption*{\label{fig:singledomain}\scriptsize{}Figure 5: Single domain clustering examples over color data (top row, {\tiny{}11}{\tiny} clusters) and simulated gaussian mixtures (second row) in 2D \& 3D respectively. $1-sigma$ final trajectory-set bandwidths have been overlaid at converged mode positions.\vspace*{1mm}} \tiny \begin{tabular*}{1\columnwidth}{@{\extracolsep{\fill}}p{17mm}|p{9mm}|p{9mm}|p{12mm}} \hline {$\mathbf{Data~\langle\#Dims,\#Classes\rangle}$}& \textbf{\emph{PRI}}& \textbf{\emph{GCE}}& \textbf{\emph{VoI}} \tabularnewline \hline $Seeds~~~~~\langle7D,3\rangle$ & ${\color{red}.89}$ \textbf{/} .86 \textbf{/} .87 & ${\color{red}.17}$ \textbf{/} .20 \textbf{/} .19 & ${\color{red}0.85}$ \textbf{/} 0.98 \textbf{/} 0.93 \tabularnewline \hline $Yeast~~~~~\langle8D,10\rangle$ & ${\color{red}.69}$ \textbf{/} .61 \textbf{/} .67 & .44 \textbf{/} ${\color{red}.39}$ \textbf{/} .47 & ${\color{red}3.03}$ \textbf{/} 3.10 \textbf{/} 3.22 \tabularnewline \hline $Letters~~~\langle16D,26\rangle$ & ${\color{red}.87}$ \textbf{/} .86 \textbf{/} .83 & .67 \textbf{/} .70 \textbf{/} ${\color{red}.62}$ & 4.96 \textbf{/} 5.16 \textbf{/} ${\color{red}4.72}$ \tabularnewline \hline \end{tabular*} \caption*{\label{tab:bsdtable}{\scriptsize{}Table 2: Results on higher dimension real world datasets from \cite{asuncion2007uci}, with a single kernel. Indicated values are of AAAMS \textbf{/} MS \textbf{/} VariableMS (\cite{bgeorgescu2003mean}) respectively, with best values in ${\color{red}red}$.}{\scriptsize}}%\\[10pt] \end{minipage} \end{figure} For image data, comparisons (\emph{Figs.}~\emph{\ref{fig:singleparamcomp}}, \emph{{\color{red}{4}}}, \emph{Table}.~\emph{{\color{red}{1}}} % \footnote{Probabilistic Rand Index (PRI), Variation of Information (VoI), Global Consistency Error (GCE), Boundary Displacement Error (BDE). The first three are clustering purity measures. PRI is a measure of the fraction of pairs of points whose labels are consistent with a given labeling. VoI and BDE are relative distance metrics between two given segmentations, based on average conditional entropy and boundary pixel difference, respectively. GCE measures the extent to which one labeling can be viewed as a refinement of the other. Higher is better for PRI while lower is better for the other three. For BSD300, the values indicate how well a segmentation corresponds to ones by human subjects. We noticed that coarser segmentatios tended to give better values. This, we suppose, was because humans tend to utilize much more comprehensive cues, and incorporate object or more holistic level semantics in their segmentations. It was noticed that PRI corresponded better to low level segment saliency than others.% }) are shown with joint domain Mean Shift implementation (JMS) from EDISON (\citep{christoudias2002synergism}), over Berkely Segmentation Dataset (\citep{martin2001database}, BSD300). BSD300 is meant for supervised algorithms - we simply clubbed the training and test images together. For sake of completeness, prior art on unsupervised image segmentation is also shown in \emph{Table}.~\emph{{\color{red}{1}}}. All indicated parameter values for AAAMS and JMS are squared. We did not search for the best performing parameter set for AAAMS, opting for a single low valued set instead. AAAMS performed significantly better than JMS, with results superior to other unsupervised image segmentation methods as well. Our experiments indicated that low base bandwidths, {\scriptsize{}$\left\langle\sigma_{base}^{r},\,\sigma_{base}^{s}\right\rangle$}{\scriptsize}, performed generally well on a good range of images (\emph{Fig}.~\emph{\ref{fig:singleparamcomp}}). This was due to the presented approach being locally adaptive and anisotropic. At similar clustering levels, AAAMS preserved more details and affected more salient segmentations. Single kernel AAAMS was tested on images and 2D, 3D gaussian mixtures at varied scales - with nice results. AAAMS results in\emph{ Figs.}~\emph{\color{red}{1(a)}, {\color{red}{5}}} are with postprocessing disabled. As\emph{ }indicated in \emph{Figs.}~\emph{\color{red}{1(a)}, {\color{red}{5}}}, reasonable local bandwidths arise, robustly identifying modes and salient clusters, by adapting according to local structure. \iffalse also sh hows two such sets with principle eigenvalues $\in[.15,1.2]$ ($\bigl\langle\sigma_{base}^{2},\,\epsilon^{2}\bigr\rangle$ was $\bigl\langle.02,\,.008\bigr\rangle$). \fi Experiments were conducted with some higher dimension datasets from \citep{asuncion2007uci} as well. \emph{Table.}~{\color{red}\emph{2}} shows initial results, along with comparisons with single domain standard Mean Shift (MS), and \citep{bgeorgescu2003mean}'s isotropic variable bandwidth implementation. Cluster count was kept the same as class count. AAAMS post-processing was disabled. \citep{bgeorgescu2003mean} first determines isotropic point bandwidths using the $k^{th}$ nearest neighbor distance heuristic, and subsequently utlizes them in single kernel mean shift iterations. Our experiments with it indicated a lack of clustering control. The datasets were meant for supervised classification, with attributes/feature components at different scales, and having uncorrelated and/or uninformative dimensions. Without any pre-processing (normalizations, component analysis) decent results were attained with a single kernel AAAMS. Note that \citep{bgeorgescu2003mean} internally normalizes the data, while AAAMS \& MS results are without any normalizations. Promising results, both qualitative and quantitative, are indicative of the efficacy of the presented approach. We intend to experiment further, especially with different merging schemes and on varied data spaces.\section{Conclusion} A generalized methodology for feature space partitioning and mode seeking was presented - leveraging synergism of adaptive, anisotropic Mean Shift and guided agglomeration. Unsupervised adaptation of full anisotropic bandwidths is useful and further enables Mean Shift clustering. We are excited about its prospects on point-normal clouds and video streams. Our experiments did indicate sparse data to be an issue. This is understandable, as it encumbers cluster growth and bandwidth development, with AAAMS behaving like conventional Mean Shift then. Future work would also focus on alleviating this issue. %------------------------------------------------------------------------ % ---------------- BIB---------------------------------------------------% List and number all bibliographical references in 9-point Times,single-spaced, at the end of your paper. \bibliographystyle{plainnat} \bibliography{egbib} %------------------------------------------------------------------------ \end{document} }

219 — 1411.4646

\caption{\label{sizeobs}Observation counterpart to Figure~\ref{flux}\textcolor{blue}{b}. Measured sizes of sample points of good quality (blue circles) and others which appear to be multiple sources (white triangles) from the \cite{2009:Yusef} sample. The sample sources of bad quality are not included. Points with white stars surrounded by blue circles represent the measurements of the six objects showed in Figure~\ref{Milky Way}.}

220 — 1411.4902

\caption{Plots of log$_{10}$ hydrogen number density for Model~A showing the fragment used in our quantitative example. The left images, a\&b, show the simulation at the age of 80~Myrs, while the right images c\&d show the simulation at the age of 90~Myrs. Images a\&c show the entire domain with the red square identifying and centered on the chosen fragment. Images b\&d are close-up images of the fragment. The red squares in panels b and d are centered on cells chosen for our example quantitative analysis in Section~\ref{subsec:example}. \textcolor{black}{A color version of this figure can be found in the on-line version of the manuscript.}}

221 — 1411.4912

\caption{ Butterfly diagram for our reference run (Run~\blue{D2} with $\Gamma_\rho=450$) at $r/R=0.75$ (top), $0.8$ (middle), and $0.95$ (bottom). }

\caption{ \blue{ Similar to the first and last panels of \Fig{fig:pa}, but the range of the color scale is adapted to the actual extrema. } }

\caption{ Upper panel: time evolution of $\bra{\BB^2}/\Beq^2(r)$ for Runs~D1 (dotted, blue), D2 (solid, black), and D3 (dashed, red). Middle panel: radial dependence of $\bra{\BB^2}/\Beq^2(r)$ at $t/\tautd=1$ for Runs~D1--D3. \blue{ The inset shows the radial dependence of $\bra{\BB^2}/\Beq^2(r)$ for Run~D2 inside the $r/R<0.8$ (helical zone) at $t/\tautd=0.2$ (dotted), $0.6$ (dashed), and $1$ (solid). } \blue{ Lower panel: radial dependence of $\bra{\BB^2}/\Beqz^2$ for fixed normalization. } %\url{256dyn_sph_sin_g1_vbc_pi} %\url{256dyn_sph_sin_g14_vbc_pi} %\url{256dyn_sph_sin_g17_vbc_pi} }

\caption{ \blue{ Time evolution of $B_r/\Beq$ at $r/R=0.98$ for a simulation %with $\sigma=0.5$ (Run~D5). %AB: should be smax with $\smax=0.5$ (Run~D5). } %\url{256dyn_sph_sin_g14_vbc_pi} }

\caption{ $B_r/\Beq$ \blue{at $r/R=0.98$} for simulations with different stratifications for Runs~Q1, Q2 and Q3 with density contrasts 2, 450 and 1400 from left to the right, respectively. %\url{256dyn_sph_sin_g1_vbc_pi} %\url{256dyn_sph_sin_g14_vbc_pi} %\url{256dyn_sph_sin_g17_vbc_pi} }

\caption{ Contours of negative (blue, solid lines) and positive (red, dashed) vertical velocity $\bra{U_r}_{kR<50}$ superimposed on a gray-scale representation of $\bra{B_r}_{kR<100}^2/\Beq^2(r)$ in Mercator projection at $r/R=0.85$ and $t/\tautd=0.7$ \blue{for Run~D1 (left panel, Run~D2 (middle panel) and, Run~D3 (right panel))}. }

222 — 1411.4958

\caption{{\bf An overview of our approach} to predicting surface normals of a scene from a single image. We separately learn top-down and bottom-up processes and use a fusion network to fuse the contradictory beliefs into a final interpretation. {\bf Top-down processes:} our network predicts a coarse $20 \times 20$ structure and a vanishing-point-aligned box layout from a set of discrete classes. {\bf Bottom-up processes:} our network predicts a structured local patch from a part of the image and line-labeling classes: \textcolor{blue}{convex-blue}, \textcolor{green}{concave-green}, and \textcolor{red}{occlusion-red}. {\bf Fusion process:} our network fuses the outputs of the two input networks, the rectified coarse normals with vanishing points(VP) and images to produce substantially better results. }

223 — 1411.5233

\caption{Topology of the power grid of the United Kingdom (UK) \cite{UK_Grid_Deloitte,Rohden12,Simonsen08} with $N=120$ synchronous machines and $M=165$ transmission lines. Generators ($P_i > 0$) are labeled \opencircle and motors ($P_i < 0$) denoted by \fullsquare. Half of the synchronous machines were chosen as generators and the other half as motors. The dashed line in the south-east denotes the transmission line $e_{\rm max}$ with the highest power flow $|K^{\rm flow}| = 1.294$ which connects machines with powers $P_{117} = -0.76$ and $P_{118} = 0.95$. After the removal of this edge the power flow in the transmission line depicted in bold (a bit to the west of the highest-flow line $e_{\rm max}$) increases most, actually it takes all the flow from $e_{\rm max}$ and it therefore defines the backup capacity $P_{\rm B} = 1.294$. Map of the UK from \cite{UK_map_url}, changed. \label{UK_grid}}

224 — 1411.5661

\caption{Interval $7$-coloring of $K_6$ and the corresponding 1-factorization $\mathfrak{F}=\left\{F_1^1, F_1^2, F_1^0, F_2^0, F_3^0\right\}$} \label{K_6factorization} \end{figure} \begin{lemma}[Equivalence lemma]\label{lEquiv} The following two statements are equivalent: \begin{description} \item{(a)} there exists $\alpha$ interval edge-coloring of $K_{2n}$ such that ${\rm sh}(\alpha) = (b_1, b_2, \ldots b_{n-1})$, \item{(b)} there exist $\mathbf{v}$ ordering of vertices and $\mathfrak{F} = \left\{F^0_j\|\j=1,2,\ldots,2n-1-\sum\limits_{i=1}^{n-1}b_i \right\}\cup \bigcup\limits_{i=1}^{n-1}\left\{F^i_j\|\j=1,2,\ldots,b_i\right\}$ 1-factorization of $K_{2n}$ such that $F^i_j$ is $i$-splitted with respect to the ordering $\mathbf{v}$, $i=1,2,\ldots,n-1$, $j=1,2,\ldots,b_i$, $b_i \in \mathbb{Z}_+$. \end{description} \end{lemma} \begin{proof} Throughout the proof we will use $B_i$ as a shorthand for $\sum\limits_{j=1}^{i}b_j$, $i=0,1,\ldots,n-1$. \begin{description} \item{(a) $=>$ (b)}. Let $\alpha$ be an interval $t$-coloring of $K_{2n}$ such that ${\rm sh}(\alpha) = (b_1, b_2, \ldots b_{n-1})$. We choose the ordering $\mathbf{v}_\alpha$ and construct the 1-factorization $\mathfrak{F}$ of $K_{2n}$. According to Remark \ref{middleColors}, there exist $2n-1-|{\rm sh}(\alpha)|$ colors that appear in the spectrums of all the vertices. By definition, $|{\rm sh}(\alpha)| = \sum\limits_{i=1}^{n-1}b_i$, so we take $F_j^0 = C_{|{\rm sh}(\alpha)|+j}(\alpha)$, for every $j=1,2,\ldots,2n-1-|{\rm sh}(\alpha)|$. For every $i=1,2,\ldots,n-1$, Remark \ref{splittedColors} implies there exist $|L_{\mathbf{v}_\alpha}^i(\alpha)|=b_i$ distinct colors that appear only in the spectrums of the first $i$ pairs of vertices and another $|R_{\mathbf{v}_\alpha}^i(\alpha)|=b_i$ distinct colors that appear only in the spectrums of the remaining $2n-2i$ vertices. We take $F^i_j = C_{B_{i-1} + j}(\alpha) \cup C_{B_{i-1} + 2n - 1 + j}(\alpha)$, for every $i=1,2,\ldots,n-1$ and $j=1,2,\ldots,b_i$. Note that the edges colored by the colors from $L_{\mathbf{v}_\alpha}^i(\alpha) \cup R_{\mathbf{v}_\alpha}^i(\alpha)$ do not cross the vertical line between the $i$-th and $(i+1)$-th pairs of vertices ($F_1^1$ and $F_1^2$ on Fig. \ref{K_6factorization}), so $F^i_j$ is $i$-splitted with respect to the ordering $\mathbf{v}_\alpha$ for all permitted $j$. \item{(b) $=>$ (a)}. Suppose $\mathfrak{F} = \left\{ F^0_j\ |\ j=1,2,\ldots,2n-1-|{\rm sh}(\alpha)| \right\} \cup \bigcup\limits_{i=1}^{n-1}\left\{F^i_j\ |\ j=1,2,\ldots,b_i\right\}$ is a 1-factorization of $K_{2n}$ with the property that $F_j^i$ is $i$-splitted perfect matching with respect to the ordering $\mathbf{v}=\left(u_1,v_1, u_2,v_2, \ldots,u_n,v_n\right)$, $i=1,2,\ldots,n-1$, $j=1,2,\ldots,b_i$. We construct $\alpha$ interval edge-coloring of $K_{2n}$ in the following way: \begin{tabular}{lll} $\alpha(e)=B_{i-1} + j$ & if $e \in l_{\mathbf{v}}^i(F_j^i)$ & $i=1,2,\ldots,n-1$, $j=1,2,\ldots,b_i$ \\ $\alpha(e)=B_{n-1} + j$ & if $e \in F_j^0$ & $j=1,2,\ldots,2n-1-B_{n-1}$\\ $\alpha(e)=B_{i-1} + 2n - 1 + j$ & if $e \in r_{\mathbf{v}}^i(F_j^i)$ & $i=1,2,\ldots,n-1$, $j=1,2,\ldots,b_i$ \end{tabular} The fact that $F^i_j$ is $i$-splitted with respect to the ordering $\mathbf{v}$ implies that every edge of $K_{2n}$ have received a color. The vertex $u_i$ (also $v_i$) is covered by all perfect matchings $F_j^0$, $j=1,2,\ldots,2n-1-B_{n-1}$, by the left parts of the matchings $F_j^{i'}$, $i'=i,i+1,\ldots,n-1$, and by the right parts of the matchings $F_j^{i'}$, $i'=1,2,\ldots,i-1$, for every $j=1,2,\ldots,b_{i'}$. So the spectrum is: \begin{align*} S(u_i, \alpha) = S(v_i, \alpha) &= \bigcup\limits_{i'=i}^{n-1}\{B_{i'-1} + j \|\j=1,2,\ldots,b_{i'}\} \\ & \cup \{B_{n-1} + j\|\j=1,2,\ldots,2n-1-B_{n-1}\} \\ & \cup \bigcup\limits_{i'=1}^{i-1}\{B_{i'-1} + 2n-1 + j \|\j=1,2,\ldots,b_{i'}\}\\ &= [B_{i-1}+1, B_{n-1}] \cup[B_{n-1}+1, 2n-1] \cup[2n, B_{i-1}+2n-1]\\ &= [B_{i-1}+1, B_{i-1}+2n-1] \end{align*} This proves that $\alpha$ is an interval $(B_{n-1} + 2n-1)$-coloring of $K_{2n}$. To complete the proof of the lemma we need to check the shift vector of the coloring $\alpha$. Note that for every $i=1,2,\ldots,n-1$, we have $\underline{S}(u_{i+1}, \alpha) - \underline{S}(u_{i}, \alpha) = B_{i}-B_{i-1} = b_i$. This shows that the ordering $\mathbf{v}_\alpha$ coincides with the ordering $\mathbf{v}$ and ${\rm sh}(\alpha) = (b_1, b_2, \ldots, b_{n-1})$. \end{description} \end{proof} \begin{remark}\label{splittedSameColor} Some of the matchings $F_j^0$ constructed in the first part of the proof of Equivalence lemma may be splitted perfect matchings as well, but for each of them both their left and right parts have the same color in the coloring $\alpha$. For example, in case $|{\rm sh}(\alpha)|=0$, $F^0_{\alpha(u_1v_1)} = C_{\alpha(u_1v_1)}(\alpha)$ is $1$-splitted perfect matching with respect to the ordering $\mathbf{v}_\alpha$. \end{remark} \begin{corollary}\label{cEquiv} For any $n\in\mathbb{N}$, $K_{2n}$ has an interval $t$-coloring if and only if it has a 1-factorization, where at least $t-2n+1$ perfect matchings are splitted. \end{corollary} \begin{proof} Construction of the desired 1-factorization from the interval $t$-coloring immediately follows from Remark \ref{totalShift} and Equivalence lemma. Remark \ref{splittedSameColor} implies that the number of the splitted perfect matchings in the obtained 1-factorization can be more than $t-2n+1$. If we have a 1-factorization of $K_{2n}$ with at least $t-2n+1$ splitted perfect matchings we can arbitrarily choose exactly $t-2n+1$ of them, then for each of them choose the $i$ for which it is $i$-splitted (the same perfect matching can be both $i$-splitted and $i'$-splitted for distinct $i$ and $i'$, the choice is again arbitrary) and apply Equivalence lemma. So, the corresponding coloring may not be uniquely determined. \end{proof} This corollary shows that finding an interval edge-coloring of $K_{2n}$ with many colors is equivalent to finding a 1-factorization with many splitted perfect matchings with respect to some ordering of vertices. For the ordering $\mathbf{v}$ we can define the maximum number of splitted perfect matchings over all 1-factorizations of $K_{2n}$. Because of the symmetry of complete graph this number does not actually depend on the chosen ordering $\mathbf{v}$, so we denote it by $\sigma_n$. \begin{theorem}[Equivalence theorem]\label{tEquiv} For every $n\in \mathbb{N}$, $W(K_{2n}) = 2n-1+\sigma_n$. \end{theorem} \section{Lower bounds} In order to obtain new lower bounds on $W(K_{2n})$ we split $K_{2n}$ into two edge-disjoint spanning regular subgraphs, find convenient 1-factorizations for each of them, and then apply Equivalence theorem for the union of these 1-factorizations. We fix the ordering of vertices of $K_{2n}$, $\mathbf{v} = \left(u_1,v_1, u_2,v_2, \ldots,u_n,v_n\right)$, and define two spanning regular subgraphs of $K_{2n}$, $K_2 \square K_n$ and $K_2 \times K_n$ (Fig. \ref{K_8products}): \begin{align*} &V(K_2 \square K_n) = V(K_2 \times K_n) = V(K_{2n})\\ &E(K_2 \square K_n) = \{u_iu_j\ |\ 1\leq i<j \leq n\} \cup \{u_iv_i\ |\ 1\leq i \leq n\} \cup \{v_iv_j\ |\ 1\leq i<j \leq n\}\\ &E(K_2 \times K_n) = \{u_iv_j\ |\ 1\leq i \neq j \leq n\} \end{align*} \begin{figure}[t!] \centering \includegraphics[width=0.43\textwidth]{K_8products.pdf} \caption{Two spanning regular subgraphs of $K_8$} \label{K_8products} \end{figure} Note that $E(K_{2n})=E(K_2 \square K_n)\cup E(K_2 \times K_n)$. We fix an ordering of vertices $\mathbf{v} = \left(u_1,v_1, u_2,v_2, \ldots,u_n,v_n\right)$ and define a special 1-factorization of $K_2 \square K_n$ which we denote by $\mathfrak{P}_n$: $\mathfrak{P}_n = \{P_0, P_1, \ldots, P_{n-1}\}$, where \begin{align*} P_0 &= \left\{\begin{tabular}{ll} $\{u_ju_{n+1-j}, v_jv_{n+1-j}\ |\ j=1,2,\ldots,\frac{n}{2}\}$ & if $n$ is even\\ $\{u_ju_{n+1-j}, v_jv_{n+1-j}\ |\ j=1,2,\ldots,\lfloor\frac{n}{2}\rfloor\} \cup \{u_{\frac{n+1}{2}}v_{\frac{n+1}{2}}\}$, & if $n$ is odd\\ \end{tabular}% \right.\end{align*} For every $i=1,2,\ldots,n-1$, $P_i = l_{\mathbf{v}}^i(P_i) \cup r_{\mathbf{v}}^i(P_i)$, where \begin{align*} l_{\mathbf{v}}^i(P_i) &= \left\{\begin{tabular}{ll} $\{u_ju_{i+1-j}, v_jv_{i+1-j}\ |\ j=1,2,\ldots,\frac{i}{2}\}$ & if $i$ is even\\ $\{u_ju_{i+1-j}, v_jv_{i+1-j}\ |\ j=1,2,\ldots,\lfloor\frac{i}{2}\rfloor\} \cup \{u_{\frac{i+1}{2}}v_{\frac{i+1}{2}}\}$, & if $i$ is odd\\ \end{tabular}% \right.\\r_{\mathbf{v}}^i(P_i) &= \left\{\begin{tabular}{ll} $\{u_{i+j}u_{n+1-j}, v_{i+j}v_{n+1-j}\ |\ j=1,2,\ldots,\frac{n-i}{2}\}$ & if $n-i$ is even\\ $\{u_{i+j}u_{n+1-j}, v_{i+j}v_{n+1-j}\ |\ j=1,2,\ldots,\lfloor\frac{n-i}{2}\rfloor\} \cup \{u_{\frac{n+i+1}{2}}v_{\frac{n+i+1}{2}}\}$, & if $n-i$ is odd\\ \end{tabular}% \right.\end{align*} \begin{figure}[t!] \centering \includegraphics[width=0.7\textwidth]{P_6.pdf} \caption{1-factorization $\mathfrak{P}_6$ of $K_2 \square K_6$} \label{P_6} \end{figure} $P_i$ is clearly an $i$-splitted perfect matching, for every $i=1,2,\ldots,n-1$. Note, that $K_2 \times K_n$ is a regular bipartite graph, so König's theorem \cite{Konig1916} implies it has a 1-factorization. If we consider the perfect matchings of any 1-factorization of $K_2 \times K_n$ as non-splitted matchings and add the perfect matchings of $\mathfrak{P}_n$ we obtain that $\sigma_n \geq n-1$. Equivalence theorem implies that this result is equivalent to Theorem \ref{tPetrosyan3n2}. In order to improve this bound we concentrate on finding a better 1-factorization of $K_2 \times K_n$. \begin{lemma}\label{l35n3} If $n \geq 2$, then $\sigma_n \geq \lfloor 1.5n \rfloor - 2$. \end{lemma} \begin{proof} We fix an ordering of vertices $\mathbf{v} = \left(u_1,v_1,u_2,v_2,\ldots,u_n,v_n\right)$ and consider two induced subgraphs: \begin{align*} G_1 &= K_2 \times K_n\left[\left\{u_1,v_1,u_2,v_2,\ldots,u_{\lfloor\frac{n}{2}\rfloor},v_{\lfloor\frac{n}{2}\rfloor}\right\}\right]\\ G_2 &= K_2 \times K_n\left[\left\{u_{\lfloor\frac{n}{2}\rfloor + 1},v_{\lfloor\frac{n}{2}\rfloor + 1}, u_{\lfloor\frac{n}{2}\rfloor + 2},v_{\lfloor\frac{n}{2}\rfloor + 2},\ldots,u_n,v_n \right\}\right] \end{align*} Both subgraphs are regular and bipartite, so according to the König's theorem \cite{Konig1916} they have 1-factorizations. Let the 1-factorizations of $G_1$ and $G_2$ be $F^l_1,F^l_2,\ldots,F^l_{\lfloor\frac{n}{2}\rfloor-1}$ and $F^r_1,F^r_2,\ldots,F^r_{\lceil\frac{n}{2}\rceil-1}$, respectively. By joining the first $\lfloor\frac{n}{2}\rfloor-1$ pairs of these matchings we form $\lfloor\frac{n}{2}\rfloor$-splitted perfect matchings of $K_2 \times K_n$ with respect to the ordering $\mathbf{v}$: \begin{center} $F_i = F^l_i \cup F^r_i$, for all $i=1,2,\ldots,\lfloor\frac{n}{2}\rfloor - 1$. \end{center} If we remove the edges $\bigcup\limits_{i=1}^{\lfloor\frac{n}{2}\rfloor-1}F_i$ from the graph $K_2 \times K_n$, the remaining graph is still a regular bipartite graph and has a 1-factorization, which we denote by $\mathfrak{F}_0$. Now, $\mathfrak{F}_0 \cup \bigcup\limits_{i=1}^{\lfloor\frac{n}{2}\rfloor-1}F_i \cup \mathfrak{P}_n$ is a 1-factorization of $K_{2n}$. The number of splitted matchings is $\lfloor\frac{n}{2}\rfloor - 1 + n - 1$. So we have $\sigma_n \geq \lfloor 1.5n \rfloor - 2$. \end{proof} By applying Equivalence theorem we obtain the following lower bound: \begin{theorem} \label{t35n3} If $n \geq 2$, then $W(K_{2n}) \geq \lfloor 3.5n \rfloor - 3$. \end{theorem} This theorem implies that $W(K_{10}) \geq 14$ which is the smallest example that disproves Conjecture \ref{conjPQ}. Next we focus on the case when $n$ is a composite number. \begin{lemma}\label{lComposite} For any $m,n \in\mathbb{N}$, $\sigma_{mn} \geq \sigma_m + \sigma_n + 2(m-1)(n-1)$. \end{lemma} \begin{proof} Let the vertex sets of $K_{2mn}$, $K_{2n}$ and $K_{2m}$ be as follows: \begin{align*} V(K_{2mn}) &= \left\{u_i^j,v_i^j\ |\ i=1,2,\ldots,n,\ j=1,2,\ldots,m\right\} \\ V(K_{2n}) &= \left\{\overline{u}_i,\overline{v}_i\ |\ i=1,2,\ldots,n \right\} \\ V(K_{2m}) &= \left\{\widetilde{u}^i,\widetilde{v}^i\ |\ i=1,2,\ldots,m \right\} \end{align*} \begin{figure}[t!] \centering \includegraphics[width=0.39\textwidth]{K_18-1.pdf} \hspace{1cm} \includegraphics[width=0.39\textwidth]{K_18-2.pdf} \caption{Several perfect matchings of $K_{18}$ constructed based on 1-factorizations $\overline{\mathfrak{F}}=\{N_1,N_2,N_1^0,N_2^0,N_3^0\}$ of $K_6$, $\widetilde{\mathfrak{F}}=\{M_1,M_2,M_1^0,M_2^0,M_3^0\}$ of $K_6$ and $\mathfrak{P}_6=\{P_0,P_1,P_2,P_3,P_4,P_5\}$ of $K_2\square K_6$ using Lemma \ref{lComposite}} \label{K_18matchings} \end{figure} We fix the following orderings of vertices of $K_{2mn}$, $K_{2n}$ and $K_{2m}$, respectively: \begin{align*} \mathbf{v} &= \left( u^1_1,v^1_1,u^1_2,v^1_2, \ldots, u^1_n,v^1_n, u^2_1,v^2_1,u^2_2,v^2_2, \ldots, u^2_n,v^2_n, \ldots, u^m_1,v^m_1,u^m_2,v^m_2, \ldots, u^m_n,v^m_n \right) \\ \overline{\mathbf{v}} &= \left( \overline{u}_1,\overline{v}_1,\overline{u}_2,\overline{v}_2, \ldots, \overline{u}_n,\overline{v}_n\right)\\ \widetilde{\mathbf{v}} &= \left( \widetilde{u}^1,\widetilde{v}^1,\widetilde{u}^2,\widetilde{v}^2, \ldots, \widetilde{u}^m,\widetilde{v}^m\right) \end{align*} Let $\overline{\mathfrak{F}} = \{ N_1,N_2,\ldots,N_{\sigma_n},N^0_1,N^0_2,\ldots, N^0_{2n-1-\sigma_n} \}$ be a 1-factorization of $K_{2n}$, where $N_i$, $i=1,2,\ldots,\sigma_n$ are splitted perfect matchings. Let $\widetilde{\mathfrak{F}} = \{ M_1,M_2,\ldots,M_{\sigma_m},M^0_1,M^0_2,\ldots, M^0_{2m-1-\sigma_m} \}$ be a 1-factorization of $K_{2m}$, where $M_i$, $i=1,2,\ldots,\sigma_m$ are splitted perfect matchings. We also need the graph $K_2 \square K_{2m}$ with the vertex set $\left\{w_i,z_i\|\i=1,2,\ldots,2m\right\}$, an ordering of its vertices $\mathbf{w} = \left( w_1,z_1,w_2,z_2,\ldots,w_{2m},z_{2m} \right)$, and its 1-factorization $\mathfrak{P}_{2m} = \{P_0,P_1,\ldots,P_{2m-1} \}$ as defined at the beginning of this section. We call the subgraph $K_2 \square K_{2m}[\{w_{2k-1},w_{2k},z_{2k-1},z_{2k}\}]$ $k$-th cell of $K_2 \square K_{2m}$, $1 \leq k \leq m$. During the proof we always assume that $x,y \in \{u,v\}$, $1 \leq s,t \leq n$ and $1 \leq p,q \leq m$. Let $\overline{\varphi}$ be a mapping which projects the edges of $K_{2mn}$ to the edges of $K_{2n}$. For every edge $x^p_sy^q_t \in E(K_{2mn})$, where $x_s \neq y_t$, we define $\overline{\varphi}(x^p_sy^q_t)=\overline{x}_s\overline{y}_t$. Next we define a mapping $\widetilde{\varphi}$ which projects the remaining edges of $K_{2mn}$ to the edges of $K_{2m}$. For every edge $x^p_sx^q_s \in E(K_{2mn})$ we define $\widetilde{\varphi}( x^p_sx^q_s ) = \widetilde{x}^p\widetilde{x}^q$. Note that the preimages $\overline{\varphi}^{-1}(\overline{e})$ for all $\overline{e} \in E(K_{2n})$ and $\widetilde{\varphi}^{-1}(\widetilde{x}^p\widetilde{x}^q)$ for all $\widetilde{x}^p\widetilde{x}^q \in E(K_{2m})$ are pairwise disjoint and their union covers the set $E(K_{2mn})$. We split the edge set $E(K_{2mn})$ into three parts the following way: \begin{align*} E(K_{2mn}) &= E^1 \cup E^2 \cup E^3 \text{, where} \\ E^1 &= \bigcup\limits_{i=1}^{\sigma_n}{\bigcup\limits_{\overline{e} \in N_i}{\overline{\varphi}^{-1}(\overline{e})}} \\ E^2 &= \bigcup\limits_{i=2}^{2n-1-\sigma_n}\bigcup\limits_{\overline{e} \in N^0_i}{\overline{\varphi}^{-1}(\overline{e})} \\ E^3 &= \bigcup\limits_{\overline{e} \in N^0_1}{\overline{\varphi}^{-1}(\overline{e})} \cup \bigcup\limits_{\widetilde{x}^p\widetilde{x}^q \in E(K_{2m})}{\widetilde{\varphi}^{-1}(\widetilde{x}^p\widetilde{x}^q)} \end{align*} The 1-factorization of $K_{2mn}$ we are going to construct is denoted by $\mathfrak{F}$ and also consists of three parts. \begin{center} $\mathfrak{F} = \mathfrak{F}^1 \cup \mathfrak{F}^2 \cup \mathfrak{F}^3$ \end{center} The set of perfect matchings $\mathfrak{F}^k$ covers the set $E^k$, $k=1,2,3$. Fig. \ref{K_18matchings} displays example perfect matchings for each of the parts in case $m=n=3$. The set $E^1$ contains the preimages of splitted perfect matchings of $K_{2n}$. To cover it, for every splitted perfect matching $N_i \in \overline{\mathfrak{F}}$, $i=1,2,\ldots,\sigma_n$, and for every perfect matching with an odd index $P_{2j+1} \in \mathfrak{P}_{2m}$, $j=0,1,\ldots,m-1$, we construct one perfect matching of $\mathfrak{F}^1$. \begin{align*} F^1_{i,j} = &F^1_{i,j,1} \cup F^1_{i,j,2} \cup F^1_{i,j,3} \cup F^1_{i,j,4}\text{, where }\\ F^1_{i,j,1} = &\bigcup\limits_{\substack{w_{2k-1}z_{2k-1} \in P_{2j+1} \\ 1 \leq k \leq m}} \left\{x_s^ky_t^k\ |\ \overline{x}_s\overline{y}_t \in l(N_i)\right\} \\ F^1_{i,j,2} = &\bigcup\limits_{\substack{w_{2k}z_{2k} \in P_{2j+1} \\ 1 \leq k \leq m}} \left\{x_s^ky_t^k\ |\ \overline{x}_s\overline{y}_t \in r(N_i)\right\} \\ F^1_{i,j,3} = &\bigcup\limits_{\substack{w_{2k-1}w_{2l-1} \in P_{2j+1} \\ 1 \leq k < l \leq m}} \left\{x_s^ky_t^l, y_t^kx_s^l\ |\ \overline{x}_s\overline{y}_t \in l(N_i)\right\} \\ F^1_{i,j,4} = &\bigcup\limits_{\substack{w_{2k}w_{2l} \in P_{2j+1} \\ 1 \leq k < l \leq m}} \left\{x_s^ky_t^l, y_t^kx_s^l\ |\ \overline{x}_s\overline{y}_t \in r(N_i)\right\}\\ \mathfrak{F}^1 = &\left\{F^1_{i,j}\ |\ i=1,2,\ldots,\sigma_n,\ j=0,1,\ldots,m-1\right\} \end{align*} For $F^1_{i,j,1}$ and $F^1_{i,j,2}$, we look for vertical edges in $P_{2j+1}$. If for some $k$, the vertical edge of the left (right) part of the $k$-th cell belongs to $P_{2j+1}$, we add the preimages of all edges of $l(N_i)$ ($r(N_i)$) in the $k$-th copy of $K_{2n}$ in $K_{2mn}$ to $F^1_{i,j,1}$ ($F^1_{i,j,2}$). Every matching $P_{2j+1}$ contains exactly two vertical edges ($w_{j+1}z_{j+1}$ and $w_{j+m+1}z_{j+m+1}$). If $m$ is odd, then one of these two belongs to the left part of its cell, and the other one belongs to the right part of its cell. If $m$ is even, then if $j$ is odd (even), both vertical edges belong to the right (left) parts of the cells. So, the number of edges in $F^1_{i,j,1}$ and $F^1_{i,j,2}$ can be calculated the following way: \begin{align*} |F^1_{i,j,1}| = &|l(N_i)|\left( (m \bmod 2)\cdot 1 + (1 - m \bmod 2)\cdot 2(1 - j \bmod 2) \right) \\ |F^1_{i,j,2}| = &|r(N_i)|\left( (m \bmod 2)\cdot 1 + (1 - m \bmod 2)\cdot 2(j \bmod 2) \right) \end{align*} For $F^1_{i,j,3}$ ($F^1_{i,j,4}$) we are looking for edges joining left side (right side) vertices of two different cells in $P_{2j+1}$. If $m$ is odd, then there are $\frac{m-1}{2}$ such edges. If $m$ is even, then there are $\frac{m}{2} - (1 - j \bmod 2)$ (in case of $F^1_{i,j,4}$: $\frac{m}{2} - (j \bmod 2)$) such edges. For every such edge which joins the $k$-th and $l$-th cells ($k<l$) we add the preimages of all edges in $l(N_i)$ ($r(N_i)$) which join the vertices in $k$-th and $l$-th copies of $K_{2n}$ in $K_{2mn}$ to $F^1_{i,j,3}$ ($F^1_{i,j,4}$). Note that for every chosen edge from $P_{2j+1}$, every edge in $l(N_i)$ ($r(N_i)$) has exactly 2 preimages in $F^1_{i,j,3}$ ($F^1_{i,j,4}$). So we have: \begin{align*} |F^1_{i,j,3}| = &2|l(N_i)|\left( (m \bmod 2)\cdot \frac{m-1}{2} + (1 - m \bmod 2)\cdot \left(\frac{m}{2} - (1 - j \bmod 2)\right) \right) \\ |F^1_{i,j,4}| = &2|r(N_i)|\left( (m \bmod 2)\cdot \frac{m-1}{2} + (1 - m \bmod 2)\cdot \left(\frac{m}{2} - (j \bmod 2)\right) \right) \end{align*} The construction of $F^1_{i,j}$ implies that it is a matching in $K_{2mn}$. To prove that it is also a perfect matching, we need to show that it has exactly $mn$ edges. \begin{align*} |F^1_{i,j}| = &|F^1_{i,j,1}| + |F^1_{i,j,2}| + |F^1_{i,j,3}| + |F^1_{i,j,4}| = \\ = &|l(N_i)|\left( (m \bmod 2)( 1 + m - 1) + (1 - m \bmod 2)\left( 2(1- j \bmod 2) + m - 2(1 - j\bmod 2) \right) \right) + \\ + &|r(N_i)|\left( (m \bmod 2)( 1 + m - 1) + (1 - m \bmod 2)\left( 2(j \bmod 2) + m - 2(j\bmod 2) \right) \right) \\ = &\left(|l(N_i)| + |r(N_i)|\right)\left((m \bmod 2)\cdot m + (1 - m \bmod 2)\cdot m \right) = nm \end{align*} The matchings $F^1_{i,j}$ and $F^1_{i',j'}$ are disjoint if $i \ne i'$ or $j \ne j'$, as their edges correspond to either different edges in $K_{2n}$ or to different edges in $K_2 \square K_{2m}$. Also note that, if $N_i$ is an $r$-splitted matching for $\overline{\mathbf{v}}$, then $F^1_{i,j}$ is $(jn+r)$-splitted matching for $\mathbf{v}$, for every $i=1,2,\ldots,\sigma_n$ and $j=0,1,\ldots,m-1$. The set $E^2$ contains the preimages of all but one non-splitted perfect matchings. To cover it, for every non-splitted perfect matching $N^0_i \in \overline{\mathfrak{F}}$ except $N^0_1$ (the choice of this exception is arbitrary) and for every perfect matching with an even index $P_{2j} \in \mathfrak{P}_{2m}$ we construct one perfect matching of $\mathfrak{F}^2$. \begin{align*} F^2_{i,j} = &F^2_{i,j,1} \cup F^2_{i,j,2}\text{, where }\\ F^2_{i,j,1} = &\bigcup\limits_{\substack{w_{2k-1}w_{2k} \in P_{2j} \\ 1 \leq k \leq m}} \left\{x_s^ky_t^k\ |\ \overline{x}_s\overline{y}_t \in N^0_i\right\} \\ F^2_{i,j,2} = &\bigcup\limits_{\substack{w_{2k-1}w_{2l} \in P_{2j} \\ 1 \leq k < l \leq m}} \left\{x_s^ky_t^l, y_t^kx_s^l\ |\ \overline{x}_s\overline{y}_t \in N^0_i\right\}\\ \mathfrak{F}^2 = &\{F^2_{i,j}\ |\ i=2,3,\ldots,2n-1-\sigma_n,\ j=0,1,\ldots,m-1\} \end{align*} The matchings $P_{2j}$ have only horizontal edges. We look for those edges which join a vertex from the left part of a cell to a vertex from the right part of a (possibly different) cell. If both endpoints of an edge belong to the same $k$-th cell, we add the preimages of all edges of $N_i^0$ which belong to the $k$-th copy of $K_{2n}$ in $K_{2mn}$ to the set $F^2_{i,j,1}$. The number of such edges in $P_{2j}$ is $1$ if $m$ is odd and $2(j \bmod 2)$ if $m$ is even. So we have: \begin{align*} |F^2_{i,j,1}| = &n\left( (m \bmod 2)\cdot 1 + (1 - m \bmod 2)\cdot 2(j \bmod 2) \right) \end{align*} If the edge of $P_{2j}$ joins vertices of $k$-th and $l$-th cells ($k < l$) then we add both preimages of all edges of $N_i^0$ which join the vertices of $k$-th and $l$-th copies of $K_{2n}$ in $K_{2mn}$ to $F^2_{i,j,2}$. The number of such edges in $P_{2j}$ is $\frac{m-1}{2}$ if $m$ is odd, and $\frac{m}{2} - (j \bmod 2)$ if $m$ is even. So, \begin{align*} |F^2_{i,j,2}| = &2n\left( (m \bmod 2)\cdot \frac{m-1}{2} + (1 - m \bmod 2)\cdot \left(\frac{m}{2} - (j \bmod 2)\right) \right) \\ |F^2_{i,j}| = &|F^2_{i,j,1}| + |F^2_{i,j,2}| =\\ = &n\left( (m \bmod 2)(1+m-1) + (1 - m \bmod 2)(2(j \bmod 2) + m - 2(j \bmod 2)) \right) =\\ = &n\left( (m \bmod 2) \cdot m + (1 - m \bmod 2) \cdot m \right) = nm \end{align*} Similar to the matchings in $\mathfrak{F}^1$, the matchings $F^2_{i,j}$ and $F^2_{i',j'}$ are disjoint if $i \ne i'$ or $j \ne j'$. Note that for every $i=2,3,\ldots,2n-1-\sigma_n$, $F^2_{i,j}$ is $jn$-splitted perfect matching for $\mathbf{v}$ for every $j=1,2,\ldots,m-1$, and is a non-splitted perfect matching if $j=0$. The set $E^3$ contains the preimages of the edges of the non-splitted perfect matching $N^0_1$ of $K_{2n}$ and the preimages of all edges of $K_{2m}[\{\widetilde{u}^1,\widetilde{u}^2,\ldots,\widetilde{u}^m\}] \cup K_{2m}[\{\widetilde{v}^1,\widetilde{v}^2,\ldots,\widetilde{v}^m\}]$. The preimages of the edges of $K_{2m}$ form $2n$ disjoint complete graphs on $m$ vertices, namely $K_{2mn}\left[\{x_s^1,x_s^2,\ldots,x_s^m\}\right]$, for every $\overline{x}_s \in V(K_{2n})$. For every edge $\overline{x}_s\overline{y}_t \in N^0_1$, its preimages together with the two copies of $K_{m}$ corresponding to the vertices $\overline{x}_s$ and $\overline{y}_t$ form the subgraph $K_{2mn}\left[\{x_s^1,y_t^1,x_s^2,y_t^2,\ldots,x_s^m,y_t^m\}\right]$, which is isomorphic to $K_{2m}$. So, the set $E^3$ consists of $n$ disjoint copies of $K_{2m}$. For every perfect matching $M \in \widetilde{\mathfrak{F}}$ we construct one perfect matching in $K_{2mn}$ by joining its $n$ disjoint copies in $E^3$: \begin{align*} F^3_{i} = & \bigcup\limits_{\substack{\overline{x}_s\overline{y}_t \in N^0_1}} \left\{ \{x_s^{p}x_s^{q}\ |\ \widetilde{u}^{p}\widetilde{u}^{q} \in M_i\} \cup \{x_s^{p}y_t^{q}\ |\ \widetilde{u}^{p}\widetilde{v}^{q} \in M_i\} \cup \{y_t^{p}y_t^{q}\ |\ \widetilde{v}^{p}\widetilde{v}^{q} \in M_i\} \right\}\\ F'^3_{i} = & \bigcup\limits_{\substack{\overline{x}_s\overline{y}_t \in N^0_1}} \left\{ \{x_s^{p}x_s^{q}\ |\ \widetilde{u}^{p}\widetilde{u}^{q} \in M^0_i\} \cup \{x_s^{p}y_t^{q}\ |\ \widetilde{u}^{p}\widetilde{v}^{q} \in M^0_i\} \cup \{y_t^{p}y_t^{q}\ |\ \widetilde{v}^{p}\widetilde{v}^{q} \in M^0_i\} \right\}\\ \mathfrak{F}^3 = &\left\{F^3_{i}\ |\ i=1,2,\ldots,\sigma_m\right\} \cup \left\{F'^3_{i}\ |\ i=1,2,\ldots,2m-1-\sigma_m\right\} \end{align*} The sets $F^3_i$ and $F'^3_i$ are pairwise disjoint matchings having $mn$ edges each. Note that if $M_i$ is $r$-splitted perfect matching for $\widetilde{\mathbf{v}}$, then $F^3_i$ is $rn$-splitted perfect matching for $\mathbf{v}$, $i=1,2,\ldots,\sigma_m$. Moreover, the perfect matchings $F'^3_i$ are not splitted, $i=1,2,\ldots,2m-1-\sigma_m$. The number of the constructed perfect matchings in $\mathfrak{F}$ is $m\sigma_n + m(2n-2-\sigma_n) + 2m-1 = 2mn-1$. Out of these the number of splitted perfect matchings is $m\sigma_n + (m-1)(2n-2-\sigma_n) + \sigma_m = \sigma_m+\sigma_n+2(m-1)(n-1)$. This completes the proof. \end{proof} By applying Equivalence theorem we obtain the following lower bound, which is a generalization of Theorem \ref{tPetrosyan4n}: \begin{theorem} \label{tComposite} For any $m,n \in\mathbb{N}$, $W(K_{2mn}) \geq W(K_{2m}) + W(K_{2n}) + 4(m-1)(n-1) - 1$. \end{theorem} We know that $W(K_6) = 7$ and $W(K_{10}) \geq 14$. The above theorem implies that $W(K_{30}) \geq 52$. This result disproves Conjecture \ref{conjLog} which predicted that $W(K_{30})=51$. But this is not the smallest case that contradicts the conjecture as we will see in Section \ref{sExact}. \begin{corollary} \label{cLower} If $n=\prod\limits_{i=1}^{\infty}p_i^{\alpha_i}$, where $p_i$ is the $i$-th prime number, $\alpha_i \in \mathbb{Z}_+$, then \begin{center} $W(K_{2n}) \geq 4n - 3 - \sum\limits_{i=1}^{\infty}{\alpha_i\left(4p_i-3-W(K_{2p_i})\right)}$. \end{center} \end{corollary} \begin{proof} Let $d_m$ denote the difference $W(K_{2m}) - (4m - 3)$. Theorem \ref{tComposite} states that $d_{mk} \geq d_m + d_k$. By induction we get $d_n \geq \sum_{i=1}^{\infty}{\alpha_i d_{p_i}}$. We complete the proof by replacing $d_{p_i}$ by its value. \end{proof} \section{Upper bounds} Let $\alpha$ be an arbitrary interval edge-coloring of $K_{2n}$, $n \in \mathbb{N}$, and $\mathbf{v}_\alpha = \left(u_1,v_1, u_2,v_2, \ldots,u_n,v_n\right)$ be its corresponding ordering of vertices. Let the shift vector of $\alpha$ be ${\rm sh}(\alpha) = (b_1,b_2,\ldots,b_{n-1})$. Equivalence lemma implies that there exists a 1-factorization of $K_{2n}$ $\mathfrak{F} = \left\{ F^0_j\ |\ j=1,2,\ldots,2n-1-\sum\limits_{i=1}^{n-1}b_i \right\} \cup \bigcup\limits_{i=1}^{n-1}\left\{F^i_j\ |\ j=1,2,\ldots,b_i\right\}$, such that $F^i_j$ is $i$-splitted with respect to the ordering $\mathbf{v_\alpha}$, $i=1,2,\ldots,n-1$, $j=1,2,\ldots,b_i$. Wherever we have an interval coloring $\alpha$ of a complete graph in the proofs of this section we will always assume that the corresponding ordering of vertices $\mathbf{v}_{\alpha}$ and 1-factorization $\mathfrak{F}$ is given. To improve the upper bounds on $W(K_{2n})$ we need several lemmas. \begin{lemma} \label{lReverse} If for some interval edge-coloring $\alpha$ of $K_{2n}$, ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$, then there exists interval edge-coloring $\beta$ of $K_{2n}$ such that ${\rm sh}(\beta) = (b_{n-1}, b_{n-2}, \ldots, b_1)$. \end{lemma} \begin{proof} Note that if some $F \in \mathfrak{F}$ is $i$-splitted with respect to $\mathbf{v}_\alpha$, then it is $(n-i)$-splitted with respect to the ordering $\mathbf{v}'_\alpha = \left(u_n,v_n,u_{n-1},v_{n-1},\ldots,u_1,v_1\right)$. We use Equivalence lemma to construct a coloring $\beta$ from $\mathfrak{F}$ with respect to the ordering $\mathbf{v}'_\alpha$. Its shift vector is $(b_{n-1}, b_{n-2}, \ldots, b_1)$. \end{proof} \begin{lemma} \label{lLessColors} If for some interval edge-coloring $\alpha$ of $K_{2n}$, ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$, where $b_i > 0$ for some $i \in [1,n-1]$, then there exists interval edge-coloring $\beta$ of $K_{2n}$ such that ${\rm sh}(\beta) = (b_1, b_2, \ldots, b_{i-1}, b_{i}-1, b_{i+1}, \ldots, b_{n-1})$. \end{lemma} \begin{proof} The condition $b_i > 0$ implies that there exists a perfect matching $F_{b_i}^i \in \mathfrak{F}$ which is $i$-splitted with respect to the ordering $\mathbf{v}_\alpha$. We construct the coloring $\beta$ by applying Equivalence lemma to the 1-factorization $\mathfrak{F}$ by regarding the perfect matching $F_{b_i}^i$ as a non-splitted one (we can rename it to $F^0_{2n-|{\rm sh}(\alpha)|}$). \end{proof} \begin{lemma} \label{l2k1} If ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$ for some interval edge-coloring $\alpha$ of $K_{2n}$, then \begin{center} $\sum\limits_{i=1}^{k}{b_i} \leq 2k-1$, for every $k=1,2,\ldots,n-1$. \end{center} \end{lemma} \begin{proof} According to the proof of Equivalence lemma, the left parts of the perfect matchings $F^i_j$ cover the vertex $u_1$ (and $v_1$), $i=1,2,\ldots,k$, $j=1,2,\ldots,b_i$. Moreover, \begin{center} $\bigcup\limits_{i=1}^{k} \bigcup\limits_{j=1}^{b_i} {l^i_{\mathbf{v}_\alpha}\left(F^i_j\right)} \subset E\left(H_{\mathbf{v}_\alpha}^{[1,k]}\right)$ \end{center} To complete the proof we note that the number of the perfect matchings $F^i_j$ is $\sum\limits_{i=1}^{k}{b_i}$ and the degree of the vertex $u_1$ (or $v_1$) in $H_{\mathbf{v}_\alpha}^{[1,k]}$ is $2k-1$. \end{proof} We will call the vector $(b_1,b_2,\ldots,b_k)$ \textit{saturated} if $\sum\limits_{i=1}^{k}{b_i} = 2k-1$. \begin{corollary} \label{c2n4} If $\alpha$ is an interval edge-coloring of $K_{2n}$, $n\geq 3$, then $|{\rm sh}(\alpha)| \leq 2n-4$. \end{corollary} \begin{proof} Let ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$. Lemma \ref{l2k1} implies that $\sum\limits_{i=1}^{n-2}{b_i} \leq 2n-5$. The same lemma in conjuction with Lemma \ref{lReverse} implies that $b_{n-1} \leq 1$. By summing these two inequalties we complete the proof. \end{proof} \begin{lemma} \label{lAfterSaturated} If ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$ for some interval edge-coloring $\alpha$ of $K_{2n}$ and $(b_1,b_2,\ldots,b_k)$ is saturated for some $k \in[2,n-2]$, then $b_{k+1} \leq 1$. \end{lemma} \begin{proof} Lemma \ref{l2k1} implies that $b_{k+1} \leq 2$. To complete the proof we need to show that $b_{k+1} \neq 2$. Suppose the contrary, $b_{k+1} = 2$. $(b_1,b_2,\ldots,b_k)$ is saturated, so the proof of Lemma \ref{l2k1} implies that the edges $u_1x_i$ and $v_1x_i$, $x \in \{u,v\}$, $i=2,3,\ldots,k$, belong to the perfect matchings $F^i_j$, $i=1,2,\ldots,k$, $j=1,2,\ldots,b_i$. Similarly, the edges $u_1u_{k+1}$, $u_1v_{k+1}$, $v_1u_{k+1}$ and $v_1v_{k+1}$ must be covered by $F^{k+1}_1$ and $F^{k+1}_2$. Now we look at the vertex $u_2$. It is covered by the left parts of the perfect matchings $F^i_j$, $i=2,3,\ldots,k$, $j=1,2,\ldots,b_i$. In total these matchings cover all but $2k-1 - \sum\limits_{i=2}^{k}{b_i} = b_1$ edges incident to $u_2$ in the subgraph $H_{\mathbf{v}_\alpha}^{[1,k]}$. Lemma \ref{l2k1} implies that $b_1 \leq 1$, so at most one edge is left uncovered. The vertex $u_2$ must be covered by the left parts of $F^{k+1}_1$ and $F^{k+1}_2$ as well. The edges $u_2u_{k+1}$ and $u_2v_{k+1}$ cannot be used as the vertices $u_{k+1}$ and $v_{k+1}$ are already covered by $F^{k+1}_1$ and $F^{k+1}_2$. Therefore, at most one edge remains for these two matchings, which is a contradiction. \end{proof} \begin{corollary} \label{c2n5} If $\alpha$ is an interval edge-coloring of $K_{2n}$, $n\geq 5$, then $|{\rm sh}(\alpha)| \leq 2n-5$. \end{corollary} \begin{proof} Let ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$. Lemma \ref{l2k1} implies that $\sum\limits_{i=1}^{n-3}{b_i} \leq 2n-7$. We consider two cases. \begin{description} \item{Case 1:} $\sum\limits_{i=1}^{n-3}{b_i} = 2n-7$. Lemma \ref{lAfterSaturated} implies that $b_{n-2} \leq 1$. Lemmas \ref{lReverse} and \ref{l2k1} imply that $b_{n-1} \leq 1$. The sum of these inequalities proves the required bound. \item{Case 2:} $\sum\limits_{i=1}^{n-3}{b_i} \leq 2n-8$. Lemmas \ref{lReverse} and \ref{l2k1} imply that $b_{n-2} + b_{n-1} \leq 3$. The sum of these inequalities completes the proof. \end{description} \end{proof} \begin{lemma} \label{lBeforeSaturated} If ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$ for some interval edge-coloring $\alpha$ of $K_{2n}$ and $(b_1,b_2,\ldots,b_k)$ is saturated for some $k \in[3,n-1]$, then $b_{k} \geq 3$. \end{lemma} \begin{proof} Suppose the contrary, $b_k \leq 2$. If $b_k=2$, then the vector $(b_1,b_2,\ldots,b_{k-1})$ is also saturated, and we obtain contradiction with Lemma \ref{lAfterSaturated}. If $b_k \leq 1$, then we have $\sum\limits_{i=1}^{k-1}{b_i} \geq 2k-2$ which contradicts Lemma \ref{l2k1}. \end{proof} \begin{lemma} \label{lEdgeCount} If ${\rm sh}(\alpha) = (b_1,b_2,\ldots, b_{n-1})$ for some interval edge-coloring $\alpha$ of $K_{2n}$ and $k \in[2,n-2]$, then \begin{center} $k(2k-1) \geq \sum\limits_{i=1}^{k}{ib_i} + \sum\limits_{i=k+1}^{\min\{2k-1,n-1\}}{(2k-i)b_i}$. \end{center} \end{lemma} \begin{proof} We consider the subgraph $H_{\mathbf{v}_\alpha}^{[1,k]}$. The number of edges in the subgraph is $k(2k-1)$. The left part of each of the perfect matchings $F^i_j$, $i=1,2,\ldots,k$, $j=1,2,\ldots,b_i$, consists of $i$ edges, and all of them belong to the subgraph $H_{\mathbf{v}_\alpha}^{[1,k]}$. The number of such edges is $\sum\limits_{i=1}^{k}{ib_i}$. Now we fix an $i \in[k+1,r]$, where $r$ denotes $\min\{2k-1,n-1\}$. The left part of each of the perfect matchings $F_j^i$, $j=1,2,\ldots,b_i$, consists of $i$ edges. At most $2i-2k$ of them can join some vertex from $H_{\mathbf{v}_\alpha}^{[1,k]}$ with some vertex from $H_{\mathbf{v}_\alpha}^{[k+1,i]}$. So at least $2k-i$ edges belong to the subgraph $H_{\mathbf{v}_\alpha}^{[1,k]}$. The number of such edges is at least $\sum\limits_{i=k+1}^{r}{(2k-i)b_i}$. \end{proof} Lemma \ref{lEdgeCount} implies that if for some fixed $k_0$ there are many $i$-splitted perfect matchings where $i \leq k_0$, then there cannot be too many $i'$-splitted perfect matchings where $k_0 < i' < \min\{2k_0-1,n-1\}$. In order to use this lemma we need to bound the sum $\sum\limits_{i=1}^{k}{ib_i}$ from below. For the numbers $k\in \mathbb{N}$ and $r \in \mathbb{Z}_+$ we define the following: \begin{align*} &T_k = \left\{\left(b_1,b_2,\ldots,b_k\right)\ |\ \exists \alpha \text{ interval coloring of }K_{2n},\ n>k,\ {\rm sh}(\alpha)=(b_1,b_2,\ldots,b_{n-1})\right\} \\ &m(k,r) = \min\limits_{(b_1,b_2,\ldots,b_{k}) \in T_k}\left\{\sum\limits_{i=1}^{k}{ib_i}\ |\ \sum\limits_{i=1}^{k}{b_i}=r\right\} \end{align*} Note that $m(k,r)$ is not defined for all pairs $(k,r)$. For example, Lemma \ref{l2k1} implies that there are no interval colorings of $K_{2n}$ for which $\sum\limits_{i=1}^{k}{b_i}=r$ if $r \geq 2k$. It is obvious that $m(1,1)=1$ and $m(k,0)=0$, $k \in \mathbb{N}$. \begin{remark} In order to calculate $m(k,r)$, $k>1$, $r>1$, it is sufficient to take the minimum over those $(b_1,b_2,\ldots,b_k) \in T_k$ for which $\sum\limits_{i=1}^{k-1}{ib_i} = m(k-1,r-b_k)$. \end{remark} Table \ref{tableMkr} lists the values of $m(k,r)$ for $k\leq 4$ and $r \leq 7$. For example, $m(3,5)$ is calculated as follows. According to the above remark the possible candidate vectors from $T_3$ are $(1,2,2)$, $(1,1,3)$, $(1,0,4)$ and $(0,0,5)$. Lemma \ref{lAfterSaturated} implies that $(1,2,2) \notin T_3$. The coloring of $K_{12}$ in Fig. \ref{fK12} proves that $(1,1,3) \in T_3$. On the other hand, sum $b_1 + 2b_2 + 3b_3$ is larger for the other two candidate vectors, so $m(3,5) = 12$. Similarly we show that $m(4,7)=20$ and the minimum is achieved on the vector $(1,2,1,3)$, which clearly belongs to $T_4$ as illustrated in the coloring of $K_{22}$ in Fig. \ref{fK22}. By applying Lemma \ref{lLessColors} to these two colorings we prove that all the other vectors listed in the Table \ref{tableMkr} belong to the corresponding $T_k$'s. \begin{figure} \centering \includegraphics[width=0.7\textwidth]{K_12-16.pdf} \caption{Interval $16$-coloring of $K_{12}$ with a shift vector $(1,1,3,0,0)$.} \label{fK12} \end{figure} \begin{table}[t!] \centering \newcolumntype{C}{>{\centering\arraybackslash}X}% \begin{tabularx}{0.8\textwidth}{c||C|C|C|C|} \backslashbox{$r$}{$k$} & $1$ & $2$ & $3$ & $4$ \\ \hline\hline $0$ & \begin{tabular}{c}$0$ \\ $(0)$ \end{tabular} & \begin{tabular}{c}$0$ \\ $(0,0)$ \end{tabular} & \begin{tabular}{c}$0$ \\ $(0,0,0)$ \end{tabular} & \begin{tabular}{c}$0$ \\ $(0,0,0,0)$ \end{tabular} \\ \hline $1$ & \begin{tabular}{c}$1$ \\ $(1)$ \end{tabular} & \begin{tabular}{c}$1$ \\ $(1,0)$ \end{tabular} & \begin{tabular}{c}$1$ \\ $(1,0,0)$ \end{tabular} & \begin{tabular}{c}$1$ \\ $(1,0,0,0)$ \end{tabular} \\ \hline $2$ & & \begin{tabular}{c}$3$ \\ $(1,1)$ \end{tabular} & \begin{tabular}{c}$3$ \\ $(1,1,0)$ \end{tabular} & \begin{tabular}{c}$3$ \\ $(1,1,0,0)$ \end{tabular} \\ \hline $3$ & & \begin{tabular}{c}$5$ \\ $(1,2)$ \end{tabular} & \begin{tabular}{c}$5$ \\ $(1,2,0)$ \end{tabular} & \begin{tabular}{c}$5$ \\ $(1,2,0,0)$ \end{tabular} \\ \hline $4$ & & & \begin{tabular}{c}$8$ \\ $(1,2,1)$ \end{tabular} & \begin{tabular}{c}$8$ \\ $(1,2,1,0)$ \end{tabular} \\ \hline $5$ & & & \begin{tabular}{c}$12$ \\ $(1,1,3)$ \end{tabular} & \begin{tabular}{c}$12$ \\ $(1,2,1,1)$ \end{tabular} \\ \hline $6$ & & & & \begin{tabular}{c}$16$ \\ $(1,2,1,2)$ \end{tabular} \\ \hline $7$ & & & & \begin{tabular}{c}$20$ \\ $(1,2,1,3)$ \end{tabular} \\ \hline \end{tabularx} \caption{ The values of $m(k,r)$. The first row of each of the cells displays the value of $m(k,r)$. The second row contains some vector $(b_1,b_2,\ldots,b_k) \in T_k$ for which $\sum\limits_{i=1}^{k}{ib_i} = m(k,r)$. } \label{tableMkr} \end{table} \begin{lemma} \label{l2n6} If $\alpha$ is an interval edge-coloring of $K_{2n}$, $n\geq 9$, then $|{\rm sh}(\alpha)| \leq 2n-6$. \end{lemma} \begin{proof} Suppose the contrary, $|{\rm sh}(\alpha)| \geq 2n-5$. Lemmas \ref{lReverse} and \ref{l2k1} imply that $\sum\limits_{i=5}^{n-1}{b_i} \leq 2n-11$. We consider three cases. \begin{description} \item{Case 1:} $\sum\limits_{i=5}^{n-1}{b_i} = 2n-11$. Lemmas \ref{lReverse}, \ref{lBeforeSaturated} and \ref{lAfterSaturated} imply that $b_{5} \geq 3$ and $b_{4} \leq 1$. We apply Lemma \ref{l2k1} for $k=3$ to show that $b_{1} + b_{2} + b_{3} = 5$ and $b_{4}=1$. Then we apply Lemma \ref{lEdgeCount} for $k=3$. The left part of the inequality is $15$. On the right side we have $\sum\limits_{i=1}^{3}{ib_i} \geq m(3,5)=12$ and $\sum\limits_{i=4}^{5}{(6-i)b_i} \geq 5$. These inequalities contradict Lemma \ref{lEdgeCount}. \item{Case 2:} $\sum\limits_{i=5}^{n-1}{b_i} = 2n-12$. Lemma \ref{l2k1} implies that $(b_1,b_2,b_3,b_4)$ is saturated. Lemma \ref{lAfterSaturated} implies that $b_5 \leq 1$. Therefore, $(b_{n-1},b_{n-2},\ldots,b_6)$ is saturated and $b_5=1$. Lemma \ref{lBeforeSaturated} implies that $b_6 \geq 3$. Now we apply Lemma \ref{lEdgeCount} for $k=4$. The left part of the inequality is $28$. On the right side, $\sum\limits_{i=1}^{4}{ib_i} \geq m(4,7)=20$ and $\sum\limits_{i=5}^{7}{(8-i)b_i} \geq 9$. These inequalities contradict Lemma \ref{lEdgeCount}. \item{Case 3:} $\sum\limits_{i=5}^{n-1}{b_i} \leq 2n-13$. Lemma \ref{l2k1} implies that $\sum\limits_{i=1}^{4}{b_i} \leq 7$. By summing these two inequalities we obtain a contradiction. \end{description} \end{proof} \begin{figure} \centering \includegraphics[width=\textwidth]{K22-37.pdf} \caption{Interval $37$-coloring of $K_{22}$ with a shift vector $(1,2,1,3,1,1,3,1,2,1)$.} \label{fK22} \end{figure} Corollaries \ref{c2n4}, \ref{c2n5}, Lemma \ref{l2n6} and Remark \ref{totalShift} imply the following upper bound on $W(K_{2n})$. \begin{theorem} \label{tUpper} If $n \geq 3$, then \begin{center} $W(K_{2n}) \leq \left\{\begin{tabular}{ll} $4n-5$, & if $n \geq 3$,\\ $4n-6$, & if $n \geq 5$,\\ $4n-7$, & if $n \geq 9$.\\ \end{tabular} \right.$ \end{center} \end{proof}\end{theorem} \section{More exact values and an improved lower bound} \label{sExact} The lower bound on $W(K_{2n})$ from Corollary \ref{cLower} depends on the values $W(K_{2p})$ where $p$ is a prime number. For $p=2$ and $p=3$ the exact values of $W(K_{2p})$ were known before \cite{Petrosyan2010}. For $p=5$ the lower bound from Theorem \ref{t35n3} coincides with the upper bound from Theorem \ref{tUpper}. The case $p=7$ is resolved by the lemma below. Finally, for the case $p=11$, the upper bound from Theorem \ref{tUpper} is achieved by the interval $37$-coloring of $K_{22}$ shown in Fig. \ref{fK22}. This coloring also rejects Conjecture \ref{conjLog}, which predicts that $W(K_{22})=36$. \begin{lemma} \label{lK14} $W(K_{14}) = 21$. \end{lemma} \begin{proof} Theorem \ref{tUpper} implies that $W(K_{14}) \leq 22$. It is sufficient to show that $K_{14}$ does not have an interval coloring with $22$ colors. Suppose the contrary, there exists $\alpha$ interval $22$-coloring of $K_{14}$. Consider its shift vector ${\rm sh}(\alpha)=(b_1,b_2,b_3,b_4,b_5,b_6)$. From Remark \ref{totalShift} we have that $\sum\limits_{i=1}^{6}{b_i}=9$. Lemma \ref{l2k1} implies that the sums of both first and last triples cannot exceed $5$. Without loss of generality we can assume that $b_1+b_2+b_3=5$ and $b_4+b_5+b_6=4$. Lemma \ref{lAfterSaturated} implies that $b_4 \leq 1$. Lemmas \ref{lReverse} and \ref{l2k1} imply that $b_5+b_6=3$ and $b_4=1$. So $b_5 \geq 2$. Now we check the inequality from Lemma \ref{lEdgeCount} for $k=3$. The left part equals $15$. On the right part we have $\sum\limits_{i=1}^{3}{ib_i} \geq m(3,5) = 12$, $\sum\limits_{i=4}^{5}{(6-i)b_i} \geq 4$. By summing these two inequalties we get a contradiction. \end{proof} The best lower bound we could obtain is the following. \begin{theorem} \label{tLower} If $n = \prod\limits_{i=1}^{\infty}{p_i^{\alpha_i}}$, where $p_i$ is the $i$-th prime number and $\alpha_i \in \mathbb{Z}_+$, then \begin{center} $W(K_{2n}) \geq 4n - 3 - \alpha_1 - 2\alpha_2 - 3\alpha_3 - 4\alpha_4 - 4\alpha_5 - \frac{1}{2}\sum\limits_{i=6}^{\infty}{\alpha_i(p_i+1)} $. \end{center} \end{theorem} \begin{proof} To prove the bound we take the bound from Corollary \ref{cLower}, set the exact values of $W(K_{2p_i})$ for the first five prime numbers and use Theorem \ref{t35n3} to bound $W(K_{2p_i})$ for $i\geq 6$, taking into account that all prime numbers except $2$ are odd. \end{proof} Table \ref{tableAll} lists obtained lower and upper bounds on $W(K_{2n})$ and all known exact values for $n \leq 18$. \begin{table}[h] \centering \begin{tabularx}{0.96\textwidth}{r||*{18}{c|}} $n$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ & $9$ & $10$ & $11$ & $12$ & $13$ & $14$ & $15$ & $16$ & $17$ & $18$ \\ \hline\hline $W(K_{2n}) \geq $ & $1$ & $4$ & $7$ & $11$ & $14$ & $18$ & $21$ & $26$ & $29$ & $33$ & $37$ & $41$ & $42$ & $46$ & $52$ & $57$ & $56$ & $64$ \\ \hline $W(K_{2n}) = $ & $1$ & $4$ & $7$ & $11$ & $14$ & $18$ & $21$ & $26$ & $29$ & $33$ & $37$ & $41$ & & & & $57$ & & \\ \hline $W(K_{2n}) \leq $ & $1$ & $4$ & $7$ & $11$ & $14$ & $18$ & $22$ & $26$ & $29$ & $33$ & $37$ & $41$ & $45$ & $49$ & $53$ & $57$ & $61$ & $65$ \end{tabularx} \caption{ Bounds on $W(K_{2n})$: The first row lists the lower bounds from Theorem \ref{tLower}, the second row lists the known exact values and the third row lists the upper bounds from Theorem \ref{tUpper}. } \label{tableAll} \end{table} \section*{Acknowledgements} We would like to thank the organizers of the 7th Cracow Conference on Graph Theory for the wonderful atmosphere. We also thank Attila Kiss for suggesting the term \textit{shift vector}. We also would like to thank the reviewers for many valuable comments. This work was made possible by a research grant from the Armenian National Science and Education Fund (ANSEF) based in New York, USA. \begin{thebibliography}{99} \bibitem{AsratianKamalian1987} A.S. Asratian, R.R. Kamalian, \textit{Interval colorings of edges of a multigraph}, Appl. Math. \textbf{5} (1987), 25--34 (in Russian). \bibitem{AsratianKamalian1994} A.S. Asratian, R.R. Kamalian, \textit{Investigation on interval edge-colorings of graphs}, J. Combin. Theory Ser. B \textbf{62} (1994) 34--43. doi:10.1006/jctb.1994.1053 \bibitem{Axenovich2002} M.A. Axenovich, \textit{On interval colorings of planar graphs}, Congr. Numer. \textbf{159} (2002), 77--94. \bibitem{GiaroKubMal2001} K. Giaro, M. Kubale, M. Malafiejski, \textit{Consecutive colorings of the edges of general graphs}, Discrete Math. \textbf{236} (2001), 131--143. doi:10.1016/S0012-365X(00)00437-4 \bibitem{Kamalian1989} R.R. Kamalian, \textit{Interval colorings of complete bipartite graphs and trees}, preprint, Comp. Cen. of Acad. Sci. of Armenian SSR, Erevan, 1989 (in Russian). \bibitem{Kamalian1990} R.R. Kamalian, \textit{Interval edge colorings of graphs}, Doctoral Thesis, Novosibirsk, 1990. \bibitem{KamalianPetrosyan2012} R.R. Kamalian, P.A. Petrosyan, \textit{A note on interval edge-colorings of graphs}, Mathematical problems of computer science \textbf{36} (2012), 13--16. \bibitem{Khachatrian2012} H. Khachatrian, \textit{Investigation on interval edge-colorings of Cartesian products of graphs}, Yerevan State University, BS thesis, 2012 (in Armenian). \bibitem{Konig1916} D. König, \textit{Uber Graphen und ihre Anwendung auf Determinantentheorie und Mengenlehre}, Math. Ann. \textbf{77} (1916), 453--465. \bibitem{Petrosyan2005} P.A. Petrosyan, \textit{Interval edge-colorings of Möbius ladders}, Proceedings of the CSIT Conference (2005), 146--149 (in Russian). \bibitem{Petrosyan2010} P.A. Petrosyan, \textit{Interval edge-colorings of complete graphs and $n$-dimensional cubes}, Discrete Math. \textbf{310} (2010), 1580--1587. doi:10.1016/j.disc.2010.02.001 \bibitem{PetrosyanKhachatrianTananyan2013} P.A. Petrosyan, H.H. Khachatrian, H.G. Tananyan, \textit{Interval edge-colorings of Cartesian products of graphs I}, Discuss. Math. Graph Theory \textbf{33} (2013), 613--632. doi:10.7151/dmgt.1693 \bibitem{Vizing1965} V.G. Vizing, \textit{The chromatic class of a multigraph}, Kibernetika \textbf{3} (1965), 29--39 (in Russian) \bibitem{West} D.B. West, \textit{Introduction to Graph Theory}, Prentice-Hall, New Jersey, 1996. \end{thebibliography} \end{document}}}\end{proof}}

225 — 1411.5950

\caption{Viscosities of PVP~360kD solutions: low-shear values from rheometry ($\bullet$); micro-rheology data obtained using 980~nm beads at $10^4\,\mathrm{Hz}$ ({\color{red}{\tiny$\blacksquare$}}); $\eta^\prime$ deduced from swimming data ({\color{forestgreen(web)}{\footnotesize$\blacktriangle$}}). Lines are best fits (see online SI). Inset: schematic showing three snapshots of a section of a flagellum (sphere, $\approx 40$~nm) cutting through a solution polymer coils ($\approx 120$~nm) (with a circular path). Coils, which are initially in the path of the flagellum section (grey), become stretched out (red), leaving a coil-sized channel of solvent}

\caption{(a)-(e) Huggins \& Kraemer representation. (red circles)$(\eta-\eta_s)/\eta_s c$ and (black squares) $\text{ln}(\eta/\eta_s)/c$ versus polymer concentration. Lines are linear fits to the data using Eq.~\ref{eq:huggins} and Eq.~\ref{eq:kraemer} simultaneously. Both quantities should be linear, and extrapolate to a unique intrinsic viscosity $[\eta]$ at $c = 0$. (a) From the PVP viscosity data of Schneider and Doetsch. Discarding the lowest-$c$ point gives $[\eta] = 1.05\pm0.02$. (b)-(e) Our PVP at four different molecular weights. (f) The scaling of intrinsic viscosity, $[\eta]$, with molecular weight, $M$, for our PVPs.}

226 — 1411.5960

\caption{\label{Fig:ener_rang}(Color online) Ratio between the variational energy per particle and the noninteracting Fermi gas energy as a function of the Jastrow term range. Results for $N=14$ ({\color{red}$\circ$}), $N=38$ ({\color{green}$\square$}), $N=54$ ({\color{blue}$\triangle$}), $N=66$ ({\color{cyan}$\triangledown$}) particles. }

\caption{\label{Fig:energy}(Color online) The energy of the system as a function of the number of particles in the box. The red empty and green filled squares are the VMC and DMC energy results, respectively. Results for the non-interacting free gas are displayed for a comparison ({\color{blue} $\ast$}). Energy in units of $E_{FG}$. }

227 — 1411.6791

\caption{{\red Control channel handshake of Real DISH.}}

228 — 1411.6847

\caption{Tracked videos of smooth-swimming {\it E.~coli} a) on plain glass, and b) inside a crystal. c) MSD on plain glass ($\circ$ = Janus; \textcolor{red}{$\square$} = {\it E.~coli}) and inside the crystal ($\times$ = Janus, \textcolor{blue}{+} = {\it E.~coli}). Solid lines: diffusive ($t$) and ballistic ($t^2$) scaling. Arrows highlight the effect of moving from glass into the crystal (g$\rightarrow$c). d)~Confocal image of a flagella stained (red) bacterium inside a colloidal crystal. Colloids (green) touch each other, but only a small, polar slice is visible. Blue: 6~s trajectory of a bacterium with shorter flagella (not shown).}

229 — 1411.7585

\caption{Projection of $\chi_{12}(t)$ after \mixtime\of mixing on the modes$m=2$, $n=1..15$ using $P^{c,s}_{12,n'm'}$ \red{as defined in equation \ref{sheareq}}. It reproduces the shear induced by the \red{differential} rotation of the galaxy.}

\caption{\red{Difference between the full color field $\chi_{12}(t=\mathrm{t_0+\mixtime})$ at time $t_0+\mixtime$, as shown in the top right panel of the figure \ref{m=2snap}, and the partial reconstruction $\chi_{12}^{\rm sh}$ defined by equation \ref{sheareq} and shown in Figure \ref{shear}. Intuitively, this field shows the change in the color field due to turbulence rather than shear.} \label{deltashear} }

230 — 1411.7749

\caption{Eigenspectra of (a) P-T and (b) SW as a function of the confining potential width for $B_{0}/J$=1. (c) First excitation energy gap of each potential. As the width increases, unbound modes become bound modes. The energy gap between ground and first excited state is maximum when there is only one bound state. (e) and (f) show the reduced adiabaticity parameter ($\mathscr{R}$) of P-T and SW respectively. Each line represents different depths of the confining potential \textcolor{blue}{$B_{0}/J=0.05$(Blue)}, \textcolor[rgb]{0.13,0.55,0.12}{$B_{0}/J=0.1$(Green)}, \textcolor{red}{$B_{0}/J=0.2$(Red)}, $B_{0}/J=0.5$, \textcolor[rgb]{0.84,0.12,0.853}{$B_{0}/J=1$(Magenta)}, \textcolor{brown}{$B_{0}/J=5$(brown)}.}

\caption{(a) Magnon energy as a function of it's position along a disordered ($\sigma_{J}$ = 0.1) chain. We used a square well guide mean $B_{0}/J=1$. Each line represent different width of potential [\textcolor{blue}{$w=3(w/a)$(blue)}, \textcolor[rgb]{0.13,0.55,0.12}{$w=6(w/a)$(green)}, \textcolor{red}{$w=10(w/a)$(red)}, $w=25(w/a)$(black), \textcolor[rgb]{0.84,0.12,0.853}{$w=50(w/a)$(magenta)}]. Increasing the potential width has an averaging effect on these fluctuations and the magnon path becomes smoother, which is helpful in high fidelity transport. (b) and (c) Standard deviation in the ground state energy ($\sigma_{gs}$) of a disordered spin chain, as potential moves across the chain for (b) SW and (c) P-T. Each line represent different $\sigma_{J}$ [\textcolor{black}{$\sigma_{J}=2\%$(Black)}, \textcolor[rgb]{0.13,0.55,0.12}{$\sigma_{J}=6\%$(Green)}, \textcolor[rgb]{0.84,0.12,0.853}{$\sigma_{J}=10\%$(Magenta)}, \textcolor{red}{$\sigma_{J}=14\%$(Red)}, \textcolor{blue}{$\sigma_{J}=18\%$(Blue)}]. Again, Increasing the spin guide width reduces the fluctuation in the ground state energy for both potentials.}

231 — 1411.7831

\caption{A schematic diagram illustrating clusters. (a) One rectangle on an even row results in reducing the maximal occupancy of two odd rows (red lines labeled \textasteriskcentered) by one. (b) Two rectangles on even rows one directly above the other results in reducing the maximal occupancy of three odd rows (red lines labeled \textasteriskcentered) by one. }

232 — 1412.0065

\caption{\footnotesize{\bf Challenges}. We contrast third person depth (a) and RGB images (b) (overlaid with pose estimates from \cite{Qian0WT014}) with depth image and RGB images (c,d) from egocentric views of daily activities. {\color{red} Hands leaving the field-of-view}, {\color{magenta} self-occlusions}, {\color{green} occlusions due to objects } and {\color{blue} malsegmentability due to interactions with the environment} are common hard cases in egocentric settings.}

233 — 1412.0626

\caption{Cross power spectrum of the reconstructed lensing convergence map from ACTPol data with a map of the CIB as measured by \planck\at 545~GHz. The power spectra from the combination of the TT, TE, EE, and EB estimators have been coadded for the D1, D5 and D6 sky regions. The errors for each patch and estimator are determined from the cross power of 2048 simulated reconstructed lensing convergence maps with the appropriate\planck\545~GHz CIB map, and neighboring errors are less than 5\% correlated. The detection significance of this lensing signal is $\AllpcSig \sigma$. The green curve shown is not a fit to these data, but rather to the \citet{planck_ciblensing/2013} data. We find a best-fit amplitude of $A = \AllpcBestfit$, with a chi-square statistic of $\AllpcChisquare$ for \dofsAllPatches\degrees of freedom, and a probability to exceed the observed chi-square of\AllpcPte. \vspace{3mm}}

\caption{Same as Figure \ref{fig:plotAllPatches_Allpc}, using only the EE and EB lensing estimators. Measurements from the D1, D5, and D6 patches are combined here as well. The polarization lensing signal is detected at a significance of $\PolonlySig \sigma$. Here we find a best-fit amplitude of $A = \PolonlyBestfit$, with a chi-square statistic of $\PolonlyChisquare$ for \dofsAllPatches\degrees of freedom, and a probability to exceed the observed chi-square of\PolonlyPte. \vspace{3mm}}

\caption{Same as Figure \ref{fig:plotAllPatches_Polonly}, using only the EB lensing estimator. The B-mode lensing signal is detected at a significance of $\EBSig \sigma$. Here we find a best-fit amplitude of $A = \EBBestfit$, with a chi-square statistic of $\EBChisquare$ for \dofsAllPatches\degrees of freedom and a probability to exceed the observed chi-square of\EBPte. \vspace{3mm}}

\caption{Comparison with other surveys. The left panel shows the temperature bandpowers from this work (red) together with those from the \planck\lensing reconstruction cross-correlated with the\planck\CIB maps at 545~GHz (purple), and the SPT lensing maps correlated with flux maps from\herschel at 500$\mu$m. The right panel shows polarization results, with the results from this work (red), the \citet[][with EE and EB estimators combined]{pbear-herschel/2013} and \citet[][SPTpol, EB only]{hanson/etal/2013}. All \herschel\results have been color-corrected by a factor of\colorscaling\to compare them to\planck\CIB results which are at a different frequency. The green solid curve is as in Figs.1--3. The dotted green curve shows the prediction using the linear matter power spectrum.}

\caption{Amplitude comparison of cross-correlation of CMB lensing with CIB emission at 545 GHz (\planck) and 500\,$\mu$m (600 GHz; \herschel) from different experiments. Shown are fits to bandpowers from Figure~\ref{fig:comparison}. We only treat statistical uncertainties in this plot. All \herschel~results have been scaled downwards by a factor of \colorscaling.}

234 — 1412.0634

\caption{\label{pdsummary}(color online) Measured (points with error bars) and fit (solid lines) Be$^+$ exponential decay plotted on a logarithmic vertical scale. Each of the three data sets consists of one experimental run, and vertical error bars represent one standard deviation. With a saturation parameter, $s$, of 0.03 and a 3 \% duty cycle (`weak cooling'), we measure a Be$^+$ lifetime of 11.4(6) hours that is limited by chemistry with background gas molecules other than H$_2$ ({\color{blue}$\blacksquare$}). Using a $>300\times$ larger photon flux for Doppler cooling ($\bullet$), the Be$^+$ number decays with a 2.19(2)-hour time constant that is consistent with an H$_2$ background pressure of $7 \times 10^{-9}$ Pa ($5 \times 10^{-11}$ Torr). To mitigate BeH$^+$ production, we trigger 200 pulses (500 Hz repetition rate) of a 157 nm excimer laser every 2 min. ({\color{red}$\blacktriangle$}). The resulting rate of BeH$^+$ photodissociation is sufficient to increase the Be$^+$ lifetime by a factor of 4.4(3) to be roughly consistent with the `weak cooling' measurements.}

235 — 1412.0866

\caption{Same quantities as above calculated in the region excluding the (negligible) part where the density is $5\%$ off compared to the reference density, as discussed in Fig~\ref{rho}. The \redc{numerical results in GC-AdResS and the full-atom simulation agree now within $3\%$, which is highly satisfactory}.}

\caption{ Pictorial representation of the AdResS scheme; CG indicates the coarse-grained region, HY the hybrid region where atomistic and coarse-grained forces are interpolated via a space-dependent, slowly varying, function $w(x)$ and AT the atomistic region (that is the region of interest). Top, the standard \redc{set-up} with the thermostat that acts globally on the whole system. Bottom, the ``local'' thermostat technique employed in this work. }

\caption{ \redc{Main figure: Potential energy of the subsystem only as a function of time, $W_{AT-AT}(t)$ compared to the energy associated to the interaction between subsystem and reservoir, $W_{AT-RES}(t)$; the former is at least one order of magnitude than the latter. Inset: The relative effect of the interaction between the AT region and the reservoir as a function of time : $\frac{|W_{AT-AT}(t)|-|W_{AT-RES}(t)|}{|W_{AT-AT}(t)|}$, it can be clearly seen that the contribution is, at most, of $10\%$. It must be underlined that in a test done with a much larger \redc{system, the} effect goes below $1.0 \%$.}}

\caption{ Molecular number density calculated with AdResS where the thermostat is acting only in the reservoir. Results are compared with the density obtained for an equivalent subsystem ($1.2 nm$) in a full atomistic NVE simulation. A discrepancy of about $5\%$ can be observed at the border of the AT region (vertical lines). Besides the fact that a discrepancy of $5\%$ is not dramatic, in general the rigorous application of GC-AdResS requires that part of the hybrid region contains a buffer of \redc{fully} atomistic molecules. Here we want to show that even in the worst-scenario-case\redc{,} the numerical accuracy is still very high.}

\caption{ Oxygen-oxygen (top), oxygen-hydrogen (middle) and hydrogen-hydrogen (bottom) radial distribution functions calculated with AdResS where the thermostat is acting only in the reservoir. Such functions are compared with the results obtained for an equivalent subsystem ($1.2 nm$) in a \redc{fully} atomistic NVE simulation \redc{and with the same quantity calculated over the entire system in the \redc{fully} atomistic simulation; the agreement is highly satisfactory.}}

\caption{ Three relevant equilibrium time correlation functions for SPC/E water at room conditions calculated with GC-AdResS and for an equivalent subsystem in a \redc{fully} atomistic NVE simulation.; as before, velocity-velocity autocorrelation function, $C_{VV}(t)$, (molecular) dipole-dipole autocorrelation function, $C_{\mu\mu}(t)$, reactive flux correlation function, $k(t)$ (semilogarithmic plot). The agreement between GC-AdResS and the \redc{fully} atomistic simulation is highly satisfactory.}

\caption{Systematic convergence of $C_{VV}(t)$, $C_{\mu\mu}(t)$ and $k(t)$ (semilogarithmic plot) of GC-AdResS to the \redc{fully} atomistic NVE results calculated over the whole system.}

236 — 1412.1135

\caption{Examples where our algorithm outperforms the previous state-of-the-art. We show the top scoring detection from the baseline detector, LSDA~\cite{lsda}, with a \textcolor{red}{Red} box and label, and the top scoring detection from our method, LSDL, as a \textcolor{green}{Green} box and label. Our algorithm improves localization (ex: rabbit, lion etc), confusion with other categories (ex: miniskirt vs maillot), and confusion with co-occurring classes (ex: volleyball vs volleyball player)}

\caption{Example mined bounding boxes learned using our method. Left side shows the mined boxes after fine-tuning with images in classification settings only, and right side shows the mined boxes after fine-tuning with auxiliary strongly annotated dataset. We show top 5 mined boxes across the dataset for corresponding category. Examples with a \textcolor{green}{green} outline are categories for which our algorithm was able to correctly mine patches of the object, while the feature space with only weak label training was not able to produce correct patches. In \textcolor{yellow}{yellow} we highlight the specific example of ``tennis racket". None of the discovered patches from the original feature space correctly located the tennis racket and instead included the person as well. After incorporating the strong annotations from auxiliary tasks, our method starts discovering tennis rackets, though still has some confusion with the person playing tennis. }

237 — 1412.1353

\caption{Ordering of classes with respect to attributes in the \textit{Shoes} dataset~\cite{whittlesearch}. Cells, coloured in \colorbox{blue!25}{blue}, represent classes that were used as negative examples and the ones coloured in \colorbox{yellow!25}{yellow} represent the ones used as positive examples for the corresponding attribute. }

238 — 1412.1740

\caption{Speed-up of {\knn} testing through SCC and SHC compression. The SCC datasets are denoted with a (C) and the SHC datasets with an (H). Results where SCC/SHC matches or exceeds the accuracy of full \knn{} (up to statistical significance) are in \textcolor{blue}{blue}.}

239 — 1412.2269

\caption{Microscopic Prominence Analysis. {\bf (a)} Structural Balance Rate. For the nodes joining the network $G_{t}$ at the same time $t$, based on their degree centrality in the network $G_{t+\Delta t}$ after $\Delta t$ timestamps, we divide them into two sets {\bf Important Nodes} and {\bf Non Important Nodes} (see \textcolor{blue}{Supporting Information, 1.2} for detail). (left) In the network $G_{t}$ we extract the sub-networks of {\bf IN} and {\bf NIN} and calculate their balance rates correspondingly. We observe that the {\bf IN} sub-network has a lower balance rate than the {\bf NIN} sub-network. (right) Similarly in the network $G_{t+\Delta t}$ we extract the sub-networks of {\bf IN} and {\bf NIN}, the {\bf IN} sub-network has a larger balance rate than the {\bf NIN} sub-network. {\bf (b)} Triad Evolution Rate. In four datasets we compute the link formation probability within different kinds of triads, we call it triad evolution rate. We observe that the ``forbidden'' triad ({\it triad 2}) (Figure~\ref{fig_tpp_example}) has much higher probability to form a new link than the disconnected sub-structure {\it triad 1}.(\textcolor{blue}{Supporting Information, 2.3}). {\bf (c)} Significance of Inferring Future Degree Centrality. We consider these centrality measures and positions as predictors of future degree centrality, we show the $p$-value associated with each feature and its corresponding significance level under Wald test (\textcolor{blue}{Supporting Information, 2.3}). {\bf (d)} Position Conditional Probability. We calculated the conditional probability of position $3$ and position $4$ (see Figure~\ref{fig_tpp_example}), $Prob(3|4)$ states the probability that a node shows up in position $3$ given the condition that it is located in position $4$; $Prob(4|3)$ is the probability that a node is located in position $4$ given the condition that it is also in position $3$. We can see that nodes in position $4$ have high probability to be located in position $3$, while nodes in position $3$ have less than $0.3$ probability to occur in position $4$.}

\caption{Generalization Performance Loss in AUPR (Degree Centrality Prediction). Different from the learning task performed in single dataset, the training set is extracted from one dataset and the prediction (testing set) is made on another dataset. AUPR, area under precision-recall curve. The AUPR score is more sensitive than AUROC in reflecting the difference of prediction \cite{acm:linkprediction6}. In order to demonstrate stability of generalization, we use AUPR for the performance evaluation. The detail of {\bf NPP} and {\bf All} methods can be found in \textcolor{blue}{Supporting Information, 3}. {\bf All} method includes existing centrality measures, which is described in \textcolor{blue}{ Supporting Information, 3.1}. Each element represents the performance reduction compared with the regular learning results (i.e., training and testing on the same dataset). We can observe that the performance reductions of {\bf NPP} method are mostly less than 20\%, while the performance reduction of {\bf All} method can achieve about 60\%.}

\caption{Predict Future Degree Centrality. We solve the future degree centrality prediction problem (\textcolor{blue}{Supporting Information, 3}) using supervised learning method. The five NPP positions (Figure~\ref{fig_tpp_example}) census contributes to our {\bf NPP} method for prediction. {\bf PA} (preferential attachment) method just includes the degree centrality feature, and {\bf TC} (triadic closure) method includes the position $3$ as feature. The method labeled {\bf All} includes existing centrality measures (\textcolor{blue}{ Supporting Information, 3.1}). The supervised learning task is to predict whether a new arriving node will become a important node or a non important node (determined by its degree centrality, see \textcolor{blue}{Supporting Information, 3.2}) in future. The experiment settings are provided in \textcolor{blue}{Supporting Information, 3.2}.}

240 — 1412.2489

\caption{\label{fig:dos_tot}Valence spectrum $w_{\rm v}$ from eq.~\eqref{eq:valence_spectrum} with broadening $\gamma_{\rm v}=0.02$ eV ({\color{dred}\full}) and $\gamma_{\rm v}=0.4$ eV ({\thickfull}). The experimental data ({\color{dblue}\dotted}) taken with incident photon energy 40.8~eV (He~II line) \cite{gouder2013} are shown for comparison.}

\caption{\label{fig:dos_f}5f component of the valence-band spectrum calculated with broadening $\gamma_{\rm v}=0.4$ eV ({\thickfull}) is compared to the experiment ({\color{dblue}\dotted}) performed at incident photon energy 1487 eV (aluminum K${}_{\alpha}$ line) \cite{teterin2013}.}

\caption{\label{fig:xps}Photoemission from the plutonium 4f levels calculated as a superposition of two spectra from eq.~\eqref{eq:core_spectrum} weighted with the statistical ratio $4/3$ ({\color{dred}\thickfull}). The 4f${}_{5/2}\,$--$\,$4f${}_{7/2}$ splitting is taken from the LDA calculation (12.64 eV), the core-valence potential $U_{\rm cv}$ is set to 6~eV, and the broadening $\gamma_{\rm c}$ is adjusted to match the width of the 4f$_{7/2}$ line in the experimental spectrum (\dotted) \cite{veal1977}. The calculated and experimental spectra are aligned at the 4f$_{7/2}$ line, and a background due to secondarily scattered electrons ({\color{dred}\shortdash}) is added to the theoretical curve as described in \cite{kotani1992}.}

241 — 1412.3354

\caption{Each center ({\color[RGB]{0,0,0}\CIRCLE}) and arm ({\color[RGB]{0,0,255}\CIRCLE}) occupy one lattice site, and each arm is physically bonded to the center({\color[RGB]{0,0,255}\textbf{---}}). The interactions between arms and their nearest neighbors are denoted by $\longleftrightarrow$.}

242 — 1412.3559

\caption{\gray light curve from August 4th 2008 to December 4th 2013. The bin size corresponds to 180- (red) and 365- days (blue). The galactic and extragalactic background emission is fixed to the best-fit parameters obtained for the overall time fit.}

243 — 1412.3585

\caption{ Left: MD results for the diffusion coefficient along several isotherms (as labeled). Right: phase diagram of the system. The points highlighted in the temperature-density plane correspond to the numerically investigated state points: a red \fulltriangle\denotes a phase separating system, a blue\fulltriangledown\corresponds to a homogeneous fluid, the black\fullsquare\is the critical point. Along the gray curves the value of the diffusion coefficient$D$ is constant and decreases from $10^{-3}$ to $7.5\cdot10^{-5}$ in units of $(2\sigma_{\rm c})^2/t_0$.}

244 — 1412.4075

\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Left panel} shows the ACF of halos of masses above ${\rm M_{th}=1\times 10^{13}h^{-1}\msun}$ (open yellow squares), ``mock CMASS galaxies", and the CCF between halos of mass above $3.5\times 10^{12}\msun$ and CMASS galaxies (open red pentagons). Black solid dots and triangles are the observed quasar-CMASS galaxy CCF and CMASS galaxy ACF (shown in both left and right panels), respectively, at $z\sim 0.53$ from \citet[][]{2013Shen}. The CMASS sample of 349,608 galaxies at $z\sim 0.53$ \citep[][]{2011aWhite,2012cAnderson} is from the Baryon Oscillation Spectroscopic Survey \citep[][]{2009aSchlegel,2013aDawson}. %The characteristic halo mass of the CMASS sample has been previously found to be $\sim 10^{13} h^{-1} M_{\odot}$, which is verified here. The sample of 8198 quasars at $0.3<z<0.9$ ($\langle z\rangle\sim 0.53$) is from the DR7 \citep{2009aAbazajian} spectroscopic quasar sample from SDSS I/II \citep[][]{2010aSchneider}. {\color{burntorange}\bf Right panel} shows the model quasar-CMASS galaxy CCF at $z=0.51$ for four cases with ${\rm (m_{h,0}, M_{h,0}, DR_0) =(2\times 10^{11}\msun, 2\times 10^{13}\msun,0.5)}$ (solid red diamonds), ${\rm (2\times 10^{11}\msun, 2\times 10^{13}\msun, 1.0)}$ (solid green hexagons), ${\rm (5\times 10^{10}\msun, 2\times 10^{13}\msun, 1.0)}$ (open blue squares) and ${\rm (2\times 10^{11}\msun, 1\times 10^{13}\msun, 1.0)}$ (open yellow stars). }

245 — 1412.4181

\caption{Performance on BSDS500 as a function of the number of training examples, before calibration ({\color{blue}blue}) and after calibration ({\color{red}red}). The smallest model was trained on $5\times10^4$ examples and the largest on $4\times10^6$ examples. Training times vary from less than one minute (40 seconds data collection $+$ 6 seconds tree training) per tree for the smallest model to under 30 minutes (15-20 minutes data collection $+$ 5 minutes tree training) per tree for the largest model.}

246 — 1412.4537

\caption{Comparison of the current density between the MD simulation (solid curves) and the fluid model (dashed curves) for \(V = 20\,\mathrm{kV}\) (bottom \textcolor{blue}{blue} curves) and \(V = 30\,\mathrm{kV}\) (top \textcolor{red}{red} curves). Other parameters used were \(d = 2500\,\mathrm{nm}\) and \(\phi = 4.7\,\mathrm{eV}\).% The value of the current density is about 15\% lower %on average in the fluid model but qualitatively gives the same results as the MD simulations. }

247 — 1412.5935

\caption{ {\bf This is a reproduction of the picture, which appears in} \cite{airoldi}. {\it Energy confinement time evolution estimated in \cite{airoldi} by solving with JETTO the balance equations (completed with a transport model): ITER97L scaling (full dots - blue line), $ITER97L^{\star}$ scaling (open dots - green line) and ITER97L(P\_red) scaling (brown line)}. }

\caption{ {\it Solutions of Eq.~(\ref{tDE11}) at the three values of ($\tau_E^0, t_0$). Blue line: ($t_0, \tau_E^0$)=(0.35sec, 0.43sec) (ITER97L ), Green line: ($t_0, \tau_E^0$)=(0.35sec, 0.625sec) ($ITER97L^{\star}$) and Brown line: ($t_0, \tau_E^0$)=(0.35sec, 0.825sec)= (ITER97L(P\_red))}. }

248 — 1412.6360

\caption{\textcolor{red}{Periodic and aperiodic experimental time series. The period, if exists, is indicated with a black line.} The parameters of the electronic circuit are $n=4$, $\alpha=3.71$, $\Gamma=40$. The initial condition is in Eq.~(\ref{Fphi}). \label{fig:temporalseries} }

249 — 1412.6677

\caption{\textbf{Uplink Throughput Gains from distributed LS-CSSP,} where the inter-site distance is around 1 kilometer, 2 cells are dynamically chosen to perform the uplink LS-CSSP in the collaborative cluster with 6 cells, and 5 interfering users in the surrounding cells with 50 \% resource block (RB) utilization ratio are configured\ref{LS-CSSP_performance}\textcolor[rgb]{1.00,0.00,0.00}{\cite{UplinkCOMP}}.}

\caption{\textbf{Enhanced soft FFR and performance evaluations of LS-CRRM schemes in H-CRANs}, where H-CRAN consists of 1 ACE and 4 RRHs, the number of RUEs is 4, the number of HUEs is 8, the number of RBs is 10, the bandwidth of each RB is 15kHz, and the mean traffic arrival rate of HUEs as 3 kbits/slot\textcolor[rgb]{1.00,0.00,0.00}{\cite{LIjian_CRRM}}.}

250 — 1412.6689

\caption{The green path {\color{green}{$R_1$}} corresponds to a step forward along the actin chain and hydrolyzing one ATP while the red path, {\color{red}{$R_2$}}$=\Gamma${\color{green}{$R_1$}} corresponds to a path in which an ATP is synthesised and moving forward one step. \textbf{Ik ben niet zeker of deze figuur echt noodzakelijk is gezien $R_1$ en $R_2$ al op figuur 1 voorkomen en de naamgeving al op figuur 2 uitgelegd werd.}}

251 — 1412.6940

\caption{Plot (a): Visualization of the target shape $z=0.1\sin\pi x$. Plots (b,c): Surface plots of the scaled objective function $\mathcal{E}/\eta_\mathrm{D}$ as a function of the scaled parameters $\eta_\Gamma/\eta_\mathrm{D}$ and $\eta_\Psi/\eta_\mathrm{D}$, for (b) pinned and (c) free boundary conditions. Plots (d,e): Distributions of $\Gamma$ (\full) and $\Psi$ (\dashed) over $X\in(0,1)$, for (d) pinned and (e) free boundary conditions.}

\caption{Plots (a,b): The two target profiles used in the axisymmetric calculations. Plots (c--f): Distributions of $\Gamma$ (\full) and $\Psi$ (\dashed) for profiles 1 (c,e) and 2 (d,f), with $\eta=\eta_\Gamma/\eta_\mathrm{D}=\eta_\Psi/\eta_\mathrm{D}$. $\Psi$ is allowed to vary in plots (c,d); $\Psi$ is set to zero for plots (e,f). Inset: legend for plots (c)--(f).}

252 — 1412.7297

\caption{(Color online) The density distribution of Erd\"os-Ren\'yi (ER) random networks and scale-free (SF) networks for different average degrees and$N=1000$. {\color{black} $\vartriangle$}, {\color{blue} $\square$}, {\color{red} $\circ$} and {\color{green} $\ast$} represent the data points of density distribution for $\langle k \rangle$=2, 4, 6 and 8, respectively. All values are averaged over 10 realizations of the networks.}

\caption{(Color online) Effect of the change in the network parameters, namely, size ($N$)and the average degree ($\langle k \rangle$) on the number of the duplicates and zero eigenvalues in different model networks. All values are averaged over 10 random realizations of the networks. Note that for ER random networks, connected component could not be obtained at $\langle k \rangle$=4 for $N$ above 1000. Here the {\color{blue} $\vartriangle$}, {\color{black} $\circ$} and {\color{red} $\square $} represent the $D_c$ of the ER, SF and configuration model networks, respectively. The solid {\color{blue} $\vartriangle$}, {\color{black} $\circ$} and {\color{black} $\square$} represent the $D_c$ of the ER, SF and configuration model networks, respectively.}

253 — 1412.7306

\caption{Our lattice set-up. Those with $m_\text{ud}^*$ are obtained by the stochastic reweighting of the Dirac operator determinant from the ensemble with the higher quark mass. \textcolor{blue}{Residual mass with $^{**}$ is estimated by weighted average of $g_i$ with some threshold.} \label{tab:setup}}

254 — 1412.8231

\caption{\textcolor{blue}{Dynamic excess kurtosis: (solid lines) $C_{4}^{d}/BFI^{2}$ as a function of dimensionless time $\tau=\nu^{2}\omega_{0}t=2\pi\nu^{2}t/T_{0}$ for different values of $R$; (dashed line) locus of transient peaks ($T_{0}$ denotes the dominant wave period and $\nu$ is the spectral bandwidth). }}

\caption{Kurtosis $\mu_{4}=\left\langle \eta^{4}\right\rangle /\left\langle \eta^{2}\right\rangle ^{2}=3(C_{4}+1)$ as a function of time $t/2T_{0}$ for JONSWAP directional wave fields initially homogenous and Gaussian ($BFI=0.78$, $\mu=0.08$, $\nu=0.15$): theoretical narrowband predictions compared against simulations and experiments (\textifsymbol[ifgeo]{49}) from \cite{Onorato2009} (data digitized from Fig. 10a,b in \cite{XiaoJFM2013}). (Left) narrow directional spreading with $\sigma_{\theta}=0.04$, $R=0.03$ (see Eq. (\ref{R})) and (Right) broad directional spreading with $\sigma_{\theta}=0.07$, $R=0.1$. Narrowband theory: dynamic kurtosis $\mu_{4}^{d}=3(C_{4}^{d}+1)$ from Eq. (\ref{c4}) (thick dashed line) and total kurtosis $\mu_{4}=\mu_{4}^{d}$+$\mu_{4}^{b}$ (thick solid line), with $\mu_{4}^{b}=3(C_{4}^{b}+1)$ from Eq. (\ref{C4b}). \textcolor{blue}{Dashed horizontal lines denote Janssen \& Bidlot (2009) dynamic kurtosis maximum from Eq. (\ref{fit}) with $b=1$.} Simulation results from \cite{XiaoJFM2013}: HOS (thin solid line), BMNLS (thin dashed line). The numerical results from \cite{Toffoli2010} are also shown: BMNLS (\Circle ) and HOS (+). }

255 — 1412.8388

\caption{Results for the empirical data sets. We consider activation times for each node as a single event sequence. In panels (a)--(c), we show the IET distributions that we obtain from combining IET distributions of all node activation sequences. The dots indicate the observed IET distributions, and the crosses indicate the estimates of the IET distributions using the KM estimator. In panels (d)--(i), we bin the event sequences according to the number of events in them ({\protect\color{plotcolor1}$\times$}: $n = 3$, {\protect\color{plotcolor2}$+$}: $n = 6$, {\protect\color{plotcolor3}$\circ$}: $n \in [8, 9]$, {\protect\color{plotcolor4}$\bigtriangledown$}: $n \in [14, 25]$, {\protect\color{plotcolor5}$\square$}: $n \in [51, 150]$). We skip every other bin to make the figure easier to read), and we normalize the IETs according to the bin's mean IET. In panels (d)--(f), we show cumulative distributions of observed IETs normalized by the mean $\mu_1^\prime$ of the observed IETs. In panels (g)--(i), we show KM estimates for the cumulative IET distributions normalized by the mean $\mu_1^{\rm{KM}}$ calculated from the estimated IET distribution. The shaded regions are the 95\% confidence intervals~\protect\cite{Denby1985Shortcut}. The data sets are (a, d, g) the Eckmann et al. e-mail data~\protect\cite{Eckmann2004Entropy}, (b, e, h) POK messages~\protect\cite{Rybski2012Communication}, and (c, f, i) the Wu et al. short-message data~\protect\cite{Wu2010Evidence}. See Fig. \protect\ref{fig:data_iets_loglog} in Appendix \protect\ref{sec:altplots} for the same distributions plotted using doubly logarithmic axes. }

\caption{Results for the empirical data sets (also see Fig.~\ref{fig:data_iets}) plotted using doubly logarithmic axes. We consider activation times for each node as a single event sequence. In panels (a)--(c), we show IET distributions that we obtain by combining IET distributions of all node activation sequences. The dots indicate the observed IET distributions, and the crosses indicate the estimates of the IET distributions using the KM estimator. In panels (d)--(i), we bin the event sequences according to the number of events in them ({\protect\color{plotcolor1}$\times$}: $n = 3$, {\protect\color{plotcolor2}$+$}: $n = 6$, {\protect\color{plotcolor3}$\circ$}: $n \in [8, 9]$, {\protect\color{plotcolor4}$\bigtriangledown$}: $n \in [14, 25]$, {\protect\color{plotcolor5}$\square$}: $n \in [51, 150]$). We skip every other bin to make the figure easier to read, and we normalize the IETs according to the bin's mean IET. In panels (d)--(f), we show cumulative distributions of observed IETs normalized by the mean $\mu_1^\prime$ of observed IETs. In panels (g)--(i), we show KM estimates for the cumulative IET distributions normalized by the mean $\mu_1^{\rm{KM}}$ calculated from the estimated IET distribution. The shaded regions are the 95\% confidence intervals~\protect\cite{Denby1985Shortcut}. The data sets are (a, d, g) the Eckmann et al. e-mail data~\protect\cite{Eckmann2004Entropy}, (b, e, h) POK messages~\protect\cite{Rybski2012Communication}, and (c, f, i) the Wu et al. short-message data~\protect\cite{Wu2010Evidence}. }

\caption{Numerical calculations for a model in which we produce the event sequences using the IET distribution $p(\tau | \tau_0) = \frac{1}{\tau_0} f(\tau / \tau_0)$, where $f(\tau)=e^{-\tau}$ and the mean values $\tau_0$ are distributed such that the expected numbers of events satisfy the probability distribution $p(n) \propto n^{-2.5}$ (where $n \geq 1$). (a) Cumulative distribution of observed IETs (green dots) and a KM estimate for the cumulative distribution (blue crosses). The black curve is the theoretical estimate of Eq.~\protect\eqref{eq:iet_collapse_plaw_exp} for the real IET distribution $p(\tau ) \propto E_{\alpha - 2}(\tau / T)$, where $E_n$ is the exponential integral function \protect\cite{DLMF}. (b) Cumulative distributions of observed IETs when we bin event sequences according to the expected number of observed events $\hat{n}=T/\tau_0$ ({\protect\color{plotcolor1}$\times$}: $\hat{n} \in (2, 3]$, {\protect\color{plotcolor2}$+$}: $\hat{n} \in (5, 6]$, {\protect\color{plotcolor3}$\circ$}: $\hat{n} \in (7, 9]$, {\protect\color{plotcolor4}$\bigtriangledown$}: $\hat{n} \in (13, 25]$, {\protect\color{plotcolor5}$\square$}: $\hat{n} \in (50, 150]$). We skip every other bin to make the figure easier to read, and we divide the IETs in each bin by the mean $\tau_0$ value of the bin $\mu_1(\tau_0)$. In the inset, we show KM estimates for the cumulative distributions IETs of each bin. (c) Cumulative distributions of observed IETs when we bin event sequences according to the observed number $n$ of events. We divide the IETs in each bin by the mean observed IET value $\mu_1^\prime$ of the bin. The lines correspond to IET distributions predicted by Eq.~\protect\eqref{eq:obs_iet_exp_cum} (or to mixtures of them for bins that have event sequences with more than one $n$ value in them). In the inset, we show KM estimates for the cumulative distributions IETs of each bin and divide the IETs in each bin with $\mu_1^{\rm{KM}}$. }