\caption{(Color online) Measured specific heat plotted as $C/T$ vs. $T$. The transitions at 38 and 41~K are broadened in the polycrystal, as the specific heat for the polycrystal rises more quickly before the onset of the transition in the single crystal. The inset shows the temperature dependence of the excess entropy \red{$\Delta S_\phi = S_\mathrm{single}-S_\mathrm{poly}.$}}
\caption{(Color online) Excess specific heat \red{plotted as $C_\phi/T$ vs. $T$.} Note, the excess specific heat can be %come negative since we neglected the pinning of the collective mode in the single crystal. The lines depict \red{$C_\phi^\mathrm{fit}/T$ in the peak region, where $C_\phi^\mathrm{fit}$} is calculated from the modified Debye law, Eq.~\eqref{eq:cphi}, with the individual pinning frequencies shown above each transition.}
\caption{\label{Vergleich}Comparison of $\theta'$, $\theta''$, $|\theta|$ and $\tau$ for different momentum cutoffs, normalised to the result for $r_{\rm opt}$ ($r_{\rm mexp}$ \textcolor{green}{$\bullet$}, $r_{\rm exp}$ \textcolor{red}{$\blacksquare$}, $r_{\rm mod}$ \textcolor{blue}{$\blacktriangle$}, $r_{\rm opt}$ $\blacktriangledown$). The relative variation, for all dimensions and all observables, is very small.}
\caption{{}(a) The ratio $g_{A}/\Delta _{A}$ (black line) and, $g_{a}/\Delta _{a}$ (red line). The perturbation treatment is valid when both $% g_{A}/\Delta _{A}$ and $g_{a}/\Delta _{a}$ less than unity. (b) The effective transverse spin coupling strength induced by SOC. (c) A comparison of eigenenergies between exact (Eqs.(\protect\ref{ExactEnergy1}) and (% \protect\ref{ExactEnergy}), lines) and perturbative(Eqs.(\protect\ref% {perturbEnergy1}) \symbol{126}(\protect\ref{perturbEnergy}), scattered dots) solutions in the large detuning regime. Our perturbation treatment is valid in the unshaded region.}
\caption{\footnotesize Results of $\Delta u_V$ and $\Delta d_V$ reconstruction for GRSV2000{\bf NLO} parametrization for both symmetric (top) and broken sea (bottom) scenarios. Solid line corresponds to the reference curve (input parametrization). Dotted line is reconstructed with MJEM and criterion (\ref{criterion}) inside the accessible for measurement region ([0.023,0.6] here). Optimal values of parameters for symmetric sea scenario for $\Delta u_V$ are $\alpha_{opt}=-0.15555$, $\beta_{opt}=-0.097951$ and for $\Delta d_V$ are $\alpha_{opt}=-0.002750$, $\beta_{opt}=-0.07190$. Optimal values of parameters for broken sea scenario for $\Delta u_V$ are $\alpha_{opt}=-0.209346$, $\beta_{opt}=0.153417$ and for $\Delta d_V$ are $\alpha_{opt}=0.702699$, $\beta_{opt}=-0.293231$. }{\includegraphics[height=4cm,width=6cm]{new-j-uv-id3-sym.eps}}{\includegraphics[height=4cm,width=6cm]{new-j-dv-id3-sym.eps}}{\includegraphics[height=4cm,width=6cm]{new-j-uv-id3-br.eps}}{\includegraphics[height=4cm,width=6cm]{new-j-dv-id3-br.eps}}
\caption{\footnotesize (color online) Idealized LO testing of the integration procedures given by Eq. (\ref{summ}) and Eq. (\ref{aint}). The input parametrizations GRSV2000LO with symmetric sea scenario (top) and broken sea scenario (bottom) are used for direct calculation of asymmetries with Eq. (\ref{difaslo}). Solid line corresponds to the input parametrization. Closed circles correspond to the values of input parametrization in the middle of each bin. Broken line shows the way of integral approximation corresponding to application of Eq. (\ref{summ}). Dashed line is obtained with MJEM and application of Eq. (\ref{aint}) for the moment calculation. Dot-dashed line is obtained with MJEM and application of Eq. (\ref{summ}) for the moment calculation. Parameters $\nu_1$ and $\nu_2$ are given by Eq. (\ref{nu}) and show the quality of reconstruction for the dashed and dot-dashed lines, respectively. }{\includegraphics[height=4cm,width=6cm]{integ-test-uv-final.eps}}{\includegraphics[height=4cm,width=6cm]{integ-test-dv-final.eps}}
\caption{Second peak contribution in $\rho_{d}$. Along the line with eq.(\ref{mean_a}), we plot the approximated estimation for $\rho_{d}$. The lower solid line (\full) shows the data from eq.(\ref{ap1}) and the data from eq.(\ref{ap2}) are depicted with upper dotted curve (\dotted). }
\caption{ (Color online) Comparison of the box-covering methods introduced by Song {\it et al.}~\cite{ss} (\textcolor{blue}{$\square$}) and in this paper (\textcolor{red}{$\circ$}). The results obtained from the two box-covering methods applied to the WWW are plotted here. The two methods yield the same fractal dimension $d_B \approx$ 4.1.}
\caption{Fractal scaling analysis (left column) and mean branching number (right column) of real-world networks including fractal (a)--(d) and non-fractal (e)--(i) networks. For each network, the original network (\textcolor{red}{$\circ$}), the skeleton (\textcolor{blue}{$\triangledown$}), and a random spanning tree (\textcolor{orange}{$\triangle$}) are studied. In (a)--(d), the straight lines, drawn for guidance, have slopes of $-4.1$, $-3.5$, $-2.3$, and $-2.1$, respectively. In (e)--(h), the fit for the exponential function in the case of the original network and a power-law fit for the random spanning tree are shown. In the right panels (a$^{\prime}$)--(i$^{\prime}$), the horizontal line at 1 is drawn for reference. }
\caption{(Color online) Fractal scaling analysis (a) and mean branching number (b) for the fractal models based on the critical branching tree with only local shortcuts with $p=0.5$ and $q=0$ (\textcolor{green}{$\square$}), and with 1\% of global shortcuts with $p=0.5$ and $q=0.01$ (\textcolor{red}{$\triangle$}). The bare critical branching tree is represented by (\textcolor{blue}{$\bigcirc$}). The solid line in (a) is guideline with a slope of --3.2. The measured degree exponent is $\gamma \approx 2.4$ and system size is $N=3\times 10^5$.}
\caption{(Color online) Fractal scaling analysis (a) and (c), and mean branching number (b) and (d) for the fractal models generated from a {\it supercritical} branching tree with $\langle n \rangle=2$, dressed by shortcuts. The data are for the bare supercritical tree (\textcolor{blue}{$\circ$}) and dressed networks with $p=0.5$ and $q=0$ (\textcolor{green}{$\square$}), $p=0.5$ and $q=0.01$ (\textcolor{red}{$\triangle$}), and $p=0$ and $q=0.05$ ($\diamond$). The solid lines in (a) and (c) are guidelines with slopes of --4.2 each. The degree exponent is $\gamma=2.3$ and system size is $N=1\times 10^5$. }
\caption{(Color online) Cumulative fraction $F_c(f)$ of the vertices counted $f$ times in the cluster-growing algorithm. $F_c(f)$ follows a power law in the small $f$ region, where the slope depends on box size $\ell_C$. However, for large values of $f$, the data largely deviate from the value extrapolated from the power-law behavior. Data are presented for $\ell_C=2$ (\textcolor{red}{$\bullet$}), $\ell_C=3$ (\textcolor{blue}{$\blacksquare$}), and $\ell_C=5$ (\textcolor{orange}{$\blacktriangle$}).}
\caption{(Color online) Box mass distribution in the cluster-growing method (a) and box-covering method (b) for the WWW. Data in (a) are for $\ell_C=2$ and those in (b) are for $\ell_B=2$ (\textcolor{red}{$\bullet$}) and $\ell_B=5$ (\textcolor{blue}{$\blacksquare$}). The solid lines are guidelines with slopes of -2.2 and -1.8, respectively.}
\caption{Average box mass versus box lateral size in the cluster-growing method (a) and the box-covering method (b) for the fractal network model constructed from a {\em critical} branching tree, dressed by shortcuts with $p=0.5$ and $q=0$ (\textcolor{red}{$\bullet$}) and $p=0.5$ and $q=0.01$ (\textcolor{blue}{$\blacksquare$}). Degree exponent is $\gamma=2.3$ and system size is $N=3\times 10^5$. Solid lines in (a) have slopes of 3.3 and 2.4, respectively, and the solid line in (b) has a slope of 3.3.}
\caption{(Color online) Average box mass as a function of box size in the cluster-growing method (a) and the box-covering method (b) for the model network constructed from a {\it supercritical} branching tree, dressed by shortcuts with $p=0.5$ and $q=0$ (\textcolor{red}{$\bullet$}) and $p=0.5$ and $q=0.01$ (\textcolor{blue}{$\blacksquare$}). Degree exponent is $\gamma=2.3$, and system size $N=1\times 10^5$. The solid line in (b) has a slope of 4.0, which is drawn for guidance.}
\caption{(Color online) Box-mass distribution in the cluster-growing method (a) and box-covering method (b) for the fractal model network grown from a {\it critical} branching tree with $\gamma=2.3$ and dressed by shortcuts generated with $p=0.5$ and $q=0.001$. The data in (b) are for $\ell_B=2$ (\textcolor{red}{$\bullet$}) and $\ell_B=5$ (\textcolor{blue}{$\blacksquare$}). Their slopes are --2.3 and --1.8, respectively. The system size is $N\approx 3\times 10^5$.}
\caption{(Color online) (a) Average box mass versus box size in the cluster-growing (\textcolor{blue}{$\times$}) and box-covering (\textcolor{red}{$\blacksquare$}) methods for a bare critical branching tree with $\gamma=2.6$. Solid lines, drawn for guidance, have slopes of 2.3 (\textcolor{blue}{$\times$}) and 2.6 (\textcolor{red}{$\blacksquare$}), respectively. (b) Box-mass distribution for the bare tree of (a). Solid guidelines have slopes of --2.8 for $\ell_B=2$ (\textcolor{red}{$\bullet$}), --2.6 for $\ell_B=5$ (\textcolor{blue}{$\blacksquare$}), and --1.6 for $\ell_B=32$ (\textcolor{green}{$\blacktriangle$}). }
\caption{(Color online) Box-mass distribution in the cluster-growing method (a) and box-covering method (b) for the fractal model network grown from a supercritical branching tree with $\langle n \rangle=2$ and $\gamma=2.3$, which is dressed by shortcuts with the parameters $p=0.5$ and $q=0$. The box size in (b) is $\ell_M=2$ (\textcolor{red}{$\bullet$}) and $\ell_M=5$ (\textcolor{blue}{$\blacksquare$}). Solid lines in (b) have slopes --2.3 and --1.8, drawn for guidance. The system size is $N\approx 1\times 10^5$.}
\caption{(Color online) Plot of the average perimeter $\langle H(M_B) \rangle$ versus box mass $M_B$ for the WWW (a) and the model network with the parameters of $\gamma=2.3$, $p=0.5$, and $q=0.0$ (b). Data are for box sizes $\ell_B=2$ (\textcolor{red}{$\bullet$}), $\ell_B=3$ (\textcolor{blue}{$\blacksquare$}), $\ell_B=5$ (\textcolor{green}{$\blacktriangle$}). The solid line (drawn as reference) has a slope of 1.0 for both (a) and (b).}
\caption{ (Color online) Degree distributions of a renormalized network of the static model for $\gamma=3$ under CG. The group masses are preassigned to follow a power law $P_m(M)\sim M^{-\eta}$ with $\eta=2$(\textcolor{red}{$\circ$}), 3 (\textcolor{blue}{$\square$}), and 4(\textcolor{green}{$\triangle$}). The degree distribution of the original network is also drawn ($\bullet$). The dashed and the dotted lines with slopes of 2.1 and 3.0, respectively, are drawn as guidelines.}
\caption{(Color online) Box-mass distribution for (a) the WWW and (c) the fractal model network of Ref.~\cite{goh2006}. Data are for $\ell_B=2$ (\textcolor{red}{$\bullet$}) and $\ell_B=5$ (\textcolor{blue}{$\blacksquare$}). The solid lines are guidelines with slopes of --2.2 and --1.8, respectively, in both (a) and (c). The degree distributions of the original network (\textcolor{green}{$\blacktriangle$}) and the renormalized networks with $\ell_B=2$ (\textcolor{red}{$\bullet$}) and $\ell_B=5$ (\textcolor{blue}{$\blacksquare$}) for (b) the WWW and (d) the fractal model network. The fractal model has a system size $N\approx 3\times 10^5$. (c) and (d) are adopted from Ref.~\cite{newjournal}.}
\caption{(Color online) Plot of the average renormalized degree $\langle k^{\prime}\rangle$ versus box mass $M_B$ for (a) the WWW and (b) the fractal model network of Ref.~\cite{goh2006}. Data are for box sizes $\ell_B=2$ (\textcolor{red}{$\bullet$}), $\ell_B=3$ (\textcolor{blue}{$\blacksquare$}), and $\ell_B=5$ (\textcolor{green}{$\blacktriangle$}). The solid lines, guidelines, have slopes of 1.0 (\textcolor{red}{$\bullet$}), 0.8 (\textcolor{blue}{$\blacksquare$}), and 0.6 (\textcolor{green}{$\blacktriangle$}), respectively, for both (a) and (b). (b) is adopted from Ref.~\cite{newjournal}.}
\caption{(Color online) Renormalized degree distributions of real-world, (a)--(c) and (e)--(h), and model, (d) and (i), networks under successive CG transformations with a fixed box size $\ell_B=2$. Symbols represent the original network (\textcolor{red}{$\circ$}), and the renormalized networks after the first (\textcolor{green}{$\triangle$}) and the second (\textcolor{blue}{$\triangledown$}) iterations. The networks of (a)--(c) are fractals, and those of (d)--(i) are non-fractals. The networks of (a)--(f) are self-similar while those of (g)--(i) are not.}
\caption{(Color online) When the coauthorship network is clustered, (a) the box-mass distribution, (b) the average renormalized degree $\langle k^{\prime} \rangle$ versus group mass $M$, and (c) the degree distribution of the original (\textcolor{green}{$\blacksquare$}) and renormalized (\textcolor{red}{$\bullet$}) networks are plotted. Groups are obtained by using the clustering algorithm of Ref.~\cite{clauset}. The solid lines, drawn for reference, have slopes of (a) $-1.6$, (b) 0.7, and (c) $-1.8$.}
\caption{(Color online) Periodic %snt $1d$ %snt lattices, $AB$ ($\circ$), $AAB$ ($\diamond$), and $AABB$ ($+$): (a) critical points obtained by series expansions in comparison with %snt analytical prediction %formula %snt for the critical line $\mu_c = (\mu_A^{1-q}\,\mu_B^{q})$, $q = 1/2$ %and $n_B = 1/2$ (\protect\rule[2pt]{0.3cm}{0.1pt}) and $q = 1/3$ % and $n_B = 1/3$ (\dashline) %\protect\rule[2pt]{0.1cm}{0.1pt} \hspace{-3pt} % \protect\rule[2pt]{0.1cm}{0.1pt}) %snt for $\mu_c=0.303228$ \cite{jensen_93}; % for $\mu_c=0.303228$ \cite{jensen_93} %snt (b) critical exponent $\beta$ in comparison with series expansion value $\beta = 0.2769$ (\dashline) %(\protect\rule[2pt]{0.1cm}{0.1pt} \hspace{-3pt} \protect\rule[2pt]{0.1cm}{0.1pt}) \cite{jensen_93} for the homogeneous case.}
\caption{(Color online) The phase diagram (a) and scaling exponent $\beta(\mu_A)$, (b) and inset of (a), for disordered lattices with various degrees of disorder: $q = 0.04$ ($\circ$), $q = 0.3$ ($\triangle$) and $q = 0.5$ ($\Box$) obtained by series expansion. The lines represent the theoretical prediction according to Eq.~(\ref{e1}), $\mu_c = (\mu_A^{1-q}\,\mu_B^q)$, for $q = 0.04$ (\protect\rule[2pt]{0.3cm}{0.1pt}), $q=0.3$ (\ddashline) and $q=0.5$ ($\cdots$) with $\mu_c=0.303228$ \cite{jensen_93}. The triangle in (a) marks the region for which the MC simulations shown in Fig.~\ref{fig:MC_bdcp} have been run. The dashed lines in (b) and the inset of (a) show the value of $\beta$ for the homogeneous case, $\beta_c = 0.2769$ (\dashline) \cite{jensen_93}. The arrow in (b) marks the homogeneous critical point.}
\caption{The number of \gray\emitting active galaxies at high galactic latitudes (galactic latitude$ |b| > 30 ^{\circ}$) predicted to be seen by the {\it GLAST LAT (Large Area Telescope)} instrument. The approximate number of sources detected by the previous {\it EGRET} instrument on the {\it Cosmic Gamma-Ray Observatory} is also shown. The curve shows the predicted integral source count {\it vs.} threshold flux.\protect\cite{ss96}.}
\caption{The optical depth of the universe to \grays\from interactions with photons of the intergalactic background light and the 2.7 K cosmic background radiation for\grays\having energies up to 100 TeV. This is given for a family of redshifts from 0.03 to 5 as indicated. The solid lines are for the fast evolution model; the dashed lines are for the baseline model.\protect\cite{sms06}.}
\caption{Contrast $K$ as a function of distance $r$ from the center of the inclusion for different depths $d$. Curves represent simulation results, symbols - experimental results. Vertical line shows the inclusion boundary. Inset: Measured width of the liquid solid-boundary as a function of depth ({\color{red} $\medbullet$} - simulation data, {\color{blue} $\square$} - experiment, line is a guide for the eye).}
\caption{Inverse of the relaxation time $1 / \tau_0$ as a function of depth obtained by directly converting the contrast $K$ based on Eq.\ref{eq:k_vs_T} neglecting the static part: {\color{red} $\medbullet$} - simulation data, {\color{blue} $\bigtriangleup$} - experiment and using the correct procedure based on Eq. \ref{eq:k_mixture}: {\color{green} $\blacksquare$} - simulation and {\color{magenta} $\Diamond$} - experiment. Horizontal dashed line - actual value of the relaxation time. }
\caption{Force measurements for various tail lengths, $L$. Oscillation frequency was varied to span a range of dimensionless lengths, $\mathcal{L}$ where the dimensionless length and force are defined in equations (\ref{DimensionlessLength}) and (\ref{eq:NondimForce}) respectively. The \textcolor{red}{+} symbols correspond to $D = \unit[0.61]{mm}$ and $a_0 = \unit[0.814]{rad}$. All other data correspond to $D = \unit[0.5]{mm}$ and $a_0 = \unit[0.435]{rad}$. There are no free parameters in the comparison between experiment and theory.}
\caption{Comparison of the time-averaged energy transfer for a Bose gas with $I=N=8$ at the interaction strength $U_0/J_0=30$. The different lines show the result for the full basis \linemediumsolid\and for truncated bases with$E_{\text{trunc}}/J_0=90$ \linemediumdashed, $E_{\text{trunc}}/J_0=60$ \linemediumdashdot, and $E_{\text{trunc}}/J_0=30$ \linemediumdotted. Depicted are the main resonance at $\omega=U_0$ (a), the non-linear resonance at $\omega=U_0/2$ (b), and the $3U_0$-resonance (c).}
\caption{Energy transfer during lattice modulation ($F=0.1$) for a superlattice system with $I=N=10$ and interaction strength~$U_0/J_0=30$ for two superlattice amplitudes~$\Delta<U_0$ in the homogenous Mott-insulator phase. The line plots show the time-averaged energy transfer as function of $\omega$ for the superlattices (A) \linemediumsolid\and (B)\linemediumsolid[lgray]\defined in figure\ref{car_bg}. The (red) arrows mark the excitation energies predicted by linear response analysis. The density plots illustrate the energy transfer as a function of frequency and time for the superlattice (A).}
\caption{Energy transfer during lattice modulation ($F=0.1$) for a superlattice system with $I=N=10$ and interaction strength~$U_0/J_0=30$ for three superlattice amplitudes~$\Delta>U_0$ in the quasi Bose-glass phase. The line plots show the time-averaged energy transfer as function of $\omega$ for the superlattices (A) \linemediumsolid\and (B)\linemediumsolid[lgray]\defined in figure\ref{car_bg}. The (red) arrows mark the excitation energies predicted by linear response analysis. The density plots illustrate the energy transfer as a function of frequency and time for the superlattice (A).}
\caption{\label{GaugeIndep} Universal eigenvalues a) $\theta'$ and b) $\theta''$ for various gauge fixing parameter and cutoffs, with $n=0,2,7$ extra dimensions. The five sets of data obtain from $r_{\rm mexp}$ \textcolor{green}{$\bullet$}, $r_{\rm exp}$ \textcolor{red}{$\blacksquare$}, $r_{\rm mod}$\textcolor{blue}{$\blacktriangle$}, $r_{\rm pow}$ \textcolor{greygreen}{$\blacklozenge$} and $r_{\rm opt}\, \blacktriangledown$.}
\caption{ (colour online) Experimental result for two equi-spaced segment apertures of angular width $2\pi/9$. Panels A, B and C show the observed angular momentum spectra for $\delta\Phi = 0$, $\pi/2$ and $\pi$, respectively, with the predicted envelope function (solid line) and experimentally measured data (bars). Even sidebands are plotted in red and odd sidebands in cyan. The main panel shows the sums of coefficients of even (red {\color{my_red} $\blacksquare$}) and odd (cyan {\color{my_cyan} $\blacksquare$}) OAM sidebands as a function of phase difference, $\delta\Phi$, between the segments. Solid curves show the predicted variation while experimental data are plotted as stars and crosses. }
\caption{(colour online) Experimental result for four equi-spaced apertures of angular width $2\pi/9$. Sums of coefficients of all $4N$ (red {\color{my_red} $\blacksquare$}), $4N+1$ (green {\color{my_green} $\blacksquare$}), $4N+2$ (cyan {\color{my_cyan} $\blacksquare$}) and $4N+3$ (blue {\color{my_blue} $\blacksquare$}), where $N$ is an integer, angular momentum sidebands as a function of phase difference between the two angular apertures. Solid curves show the predicted variation.}
\caption{The range of heliospheric distances probed by the IC \grays\versus angular distance from the Sun. 90\% of the predicted \gray\flux is produced in the region between the solid and dashed lines. Dotted lines show the solar size.}
\caption{Differential intensities for selected $\theta$. Line-sets (top to bottom): $0.3^\circ$, 1$^\circ$, 5$^\circ$, 10$^\circ$, 45$^\circ$, and 180$^\circ$. Solid line: no modulation; dashed line: $\Phi_0 = 500$ MV; dotted line: $\Phi_0 = 1000$ MV. Data points: diffuse extragalactic \gray\flux\citep{Strong2004}.}
\caption{Differential intensities for selected $\theta$. Line-sets (top to bottom): $0.3^\circ$, 1$^\circ$, 5$^\circ$, 10$^\circ$, 45$^\circ$, and 180$^\circ$. Solid line: no modulation; dashed line: $\Phi_0 = 500$ MV; dotted line: $\Phi_0 = 1000$ MV. Data points: diffuse extragalactic \gray\flux (A. W. Strong, I. V. Moskalenko,\& O. Reimer[ApJ, 613,956 (2004)]). }
\caption{Schematic diagram of thermoelectric measurement apparatus. The sample was fixed by a Apezion\textregistered{} high vacuum grease to the copper heating elements to ensure the thermal contact. The thermocouples were cemented with Stycast\textregistered{} on the outer gold pads. The electromotive force $\Delta V$ was measured on the inner gold pads. The gold wires were glued with standard colloidal silver on the gold contacts. Thermocouples $T_1$, $T_2$ and $T_3$ were used to measure the temperature of the heating element and of the copper mass (cryostat). The cryostat was electrically insulated and normally grounded ensuring $V_G = 0\, \mathrm{V}$.}
\caption{Three power spectral densities $f_1$ ({\color{red}--}), $f_2$ ({\color{green}- -}), $f_3$ ({\color{blue}-$\cdot$-})}
\caption{\label{fig:tcvsx}Bulk critical temperature $T\ped{c}$ measured as a function of the aluminum content $x$. \fullcircle: Data obtained in our single crystals by DC susceptibility \cite{karpinski05}. \opentriangle: Data from \cite{klein06}, obtained in single crystals by specific-heat measurements. \opensquare: Data from DC susceptibility in long-annealed polycrystals \cite{zambano05}.}
\caption{(a) Energy gap amplitudes $\Delta\ped{\sigma}$ (\opencircle) and $\Delta\ped{\pi}$ (\fullcircle) measured by PCAR in single crystals, as a function of the Al content determined by EDX. Error bars indicate the uncertainty on the gap values for a given curve. (b) The same values as a function of the critical temperature of the contact $T\ped{c}\apex{A}$. The gaps (obtained by PCAR) from refs. \cite{putti05}(squares), \cite{klein06}(triangles) and \cite{szabo06} (down triangles) are also reported for comparison. The dashed lines are only guides to the eye, while the straight grey line in (b) indicates the BCS $\Delta$ vs. $T_c$ dependence.}
\caption{\label{fig:parameters}(a): normalized experimental conductance curve of a point contact in a crystal with $x=0.32$ (\opencircle) compared to the single-band (\dashed) and two-band (\full) BTK fit. (b) The values of $\Delta$ (\fullcircle), $\Gamma$ (\opentriangle) and $Z$ (\opensquare) for the $\pi$ band, that allow fitting the same curve, in a series of different fits. (c) The same as in (b) but for the $\sigma$ band. The variations in the gaps are qualitatively indicated by grey strips.}
\caption{(color online) Current as a function of bias voltage $V$ at a few values of gate voltage %\textcolor[rgb]{0.00,0.00,0.63} %\textcolor[rgb]{0.00,0.00,0.63}{$\bullet$} $V_g=3.214$ V (\textcolor[rgb]{0.00,0.00,0.63}{$\bullet$}); %\textcolor[rgb]{1.00,0.00,0.00} $V_g=3.220$ V (\textcolor[rgb]{1.00,0.00,0.00}{$\circ$}); %\textcolor[rgb]{0.00,0.50,0.00} $V_g=3.226$ V (\textcolor[rgb]{0.00,0.50,0.00}{$\blacksquare$}); %\textcolor[rgb]{0.00,0.00,1.00} $V_g=3.232$ V (\textcolor[rgb]{0.00,0.00,1.00}{$\square$}); %\textcolor[rgb]{0.44,0.00,0.87} $V_g=3.244$ V (\textcolor[rgb]{0.44,0.00,0.87}{$\blacktriangle$}); %\textcolor[rgb]{1.00,0.00,0.50} $V_g=3.344$ V (\textcolor[rgb]{1.00,0.00,0.50}{$\triangle$}). %\textcolor[rgb]{0.62,0.62,0.00} The solid straight line displays the normal state IV-curve measured in a magnetic field of $B=0.2$ T at $V_g=3.202$ V. The arrows A and B illustrate the determination of the maximum supercurrent $I_{cm}$ in the non-hysteretic and hysteretic cases. The inset on the lower right illustrates a magnification of the low voltage part of the IV-curve at $V_g=3.214$ V where clear hysteresis is visible and the critical current $I_{cm}=0.15$ nA. The inset on the upper left displays an AFM image of our sample (scale bar: 1 $\mu$m). The data were measured in a two-lead configuration on the middle section at $T=65$ mK.}
\caption{\label{fig:1} The amplitudes of the projections of the exact eigenfunctions of the Hamiltonian (\ref{HO_Hamiltonian}) with $\omega=1$ to the detail spaces $W_m$ for $m=0,\ldots,6$. The sign $\lozenge$ stands for the ground state $i=0$, whereas the signs $\times$, $\ast$, $\circ$, $+$ and \drawbox{1ex}{1ex} denote the excitations $i=1,2,3,4,5$, respectively. Atomic units are used.}
\caption{\label{fig:x} The number of the significant basis functions in various subspaces required to reproduce the $\|P_0\Psi_i\|^2$ and $\|Q_m\Psi_i\|^2$ to a given precision. The horizontal axis shows the number of digit 9 in the threshold value $\eta$. The sign $+$ stands for the restricted scaling function subspace $\widetilde{V}_0$, whereas the signs $\circ$, \drawbox{1ex}{1ex}$\,$, $\lozenge$, $\triangledown$, $\vartriangle$, $\bigstar$ and $\ast$ denote the detail spaces $\widetilde{W}_m$, with $m=0,1,2,3,4,5,6$, respectively. The results for the ground state ($i=0$) are plotted in the left column both in linear and log scale, while the right column corresponds to the third excited state.}
\caption{\label{fig:y} The errors of the approximate energies $E_i^{[M]}$ determined in the Hilbert space $\mathcal{H}^{[M]}$. The sign $\lozenge$ stands for the ground state $i=0$, whereas the signs $\times$, $\ast$, $\circ$, $+$ and \drawbox{1ex}{1ex} denote the excitations $i=1,2,3,4,5$, respectively. Atomic units are used.}
\caption{\label{fig:Energ}The errors of the approximate energies $\widetilde{E}_i^{[M]}$ determined in the Hilbert space $\widetilde{\mathcal{H}}^{[M]}$ for the ground state $i=0$ and for an excited state $i=3$. The signs $\times$, \drawbox{1ex}{1ex}$\,$, $\lozenge$, $\ast$, $\triangle$ and $\circ$ stand for the threshold values $\eta=0.9,\ 0.99,\ 0.999,\ 0.9999,\ 0.99999$ and 0.999999, respectively. Atomic units are used.}
\caption{Size of the central peak after 20~ms of free expansion versus the number of atoms. The clouds are released from the following potentials: magnetic trap (red \textcolor{red}{$\circ$}), magnetic trap and disorder (black $\Box$), magnetic trap and optical lattice (blue \textcolor{blue}{$\lhd$}), magnetic trap and disorder and lattice (green \textcolor{green}{$\rhd$}). The lines correspond to a theoretical prediction (see text). The lattice depth is 6.5~$E_r$ and the disorder has a depth of 0.1~$E_r$.}
\caption{Growth rate spectra $\gLin(m)$ of unstable DTM eigenmodes for the $q$ profile in figure~\protect\ref{fig:equlib_q}. (\fullcircle): collisional case studied in this paper ($\SHp = 10^6$, $\RHp = 10^7$, $d_{\rm e} = 0$). For the parameter values $\SHp = 10^8$ and $\RHp = 10^7$, further spectra are shown for $d_{\rm e} = 0$, $0.005$ and $0.01$. The case with $d_{\rm e} = 0.01$ (\opencircle) is the one used in this paper to study the nonlinear evolution of collisionless DTMs. Only growth rates on the dominant eigenmode branch ($\Mb$-type, cf.~figure~\protect\ref{fig:mstruc}) are shown. The fastest growing modes are indicated by arrows.}
\caption{$d_{\rm e}$ dependence of the linear growth rate of the modes $m=2$, $6$ and $10$. The scanned range $10^{-4} \leq d_{\rm e} \leq 5\times 10^{-2}$ is roughly divided into three regimes: predominantly collisional, collisionless, and a regime where the skin depth $d_{\rm e}$ becomes comparable to the inter-resonance distance $D_{12}$. Both eigenmode branches $\Ma$ (\broken) and $\Mb$ (\dotted) are shown (cf.~figure~\protect\ref{fig:mstruc}).}
\caption{(color online) The partial wave phase shifts for collisions between the $|1\rangle$ and $|2\rangle$ states. The symbols represent the {\it s} (\textcolor{blue}{$\triangle$}), {\it p} (\textcolor{mygreen}{$\bigcirc$}) and {\it d} (\textcolor{red}{$\Box$}) phase shifts extracted from the data and the solid lines are a theoretical calculation from a coupled-channels model.}
\caption{\footnotesize We depict the staircase pattern for the foliation of $\ta$ and color code the four graphs of its foliation: positive vertices are labeled $+$ and are coded \textcolor{LRed}{red}; negative vertices are labeled $-$ and coded \textcolor{LBlue}{blue}; edges in $G_{+,+}$ are coded \textcolor{LRed}{red}; edges in $G_{-,-}$ are coded \textcolor{LBlue}{blue}; edges in $G_{-,+}$ are coded \textcolor{LMagenta}{light magenta}; and edges in $G_{+,-}$ are coded \textcolor{LGreen}{green}. We will always depict the meridian curves as \textcolor{DMagenta}{dark magenta} and $X$ as \textcolor{black}{black}. Thus, are color code for the graphs is: \textcolor{LRed}{$G_{+,+}$}; \textcolor{LBlue}{$G_{-,-}$}; \textcolor{LGreen}{$G_{+,-}$}; and \textcolor{LMagenta}{$G_{-,+}$}. The $\bb$-support, $\cS_K$ is colored \textcolor{LOrange}{light orange}.}
\caption{Temporal evolution of local gain at different intensities at $T = 297 K $ \textcolor[rgb]{0.98,0.00,0.00}{($e_{n}^{th}=16.31\second^{-1}$ calculated from equation \ref{toutou})},~~$\Lambda = 5~ \micro\metre$,~~$E_{0} =10~\kilo\volt\per\centi\metre$, crystal thickness $L=12mm $ : (a) $I_{0}=25.5~\milli\watt\per\centi\metre\squared \sim Ires$, (b) $I_{0}=15~\milli\watt\per\centi\metre\squared$,~~(c)$I_{0}=50~\milli\watt\per\centi\metre\squared$. ~$\tau _{r}$: characteristic time constant of amplification.}
\caption{\textcolor[rgb]{0.98,0.00,0.00}{Characteristic time of amplification versus temperature at three different intensities. The time constant at the resonance intensity is given for each temperature by the dotted line. }}
\caption{Measured (a) $Q\msub{i}$ and (b) $\Delta\lambda$ for the (\textcolor{magenta}{$+$},\textcolor{magenta}{$\times$}) GaAs TE- and TM-polarized microdisk modes and (\textcolor{cyan}{{\footnotesize $\bigcirc$}},\textcolor{cyan}{$\square$}) Al$_{0.18}$Ga$_{0.82}$As TE and TM modes, respectively. In (c) and (d), connected points represent calculated bounds on (c) $Q\msub{ss}$ and (d) $Q\msub{a}$ for each family of modes. The data were compiled from two disks of each material.}
\caption[]{[a] Intensities of C, B, Be, and Li shown for two levels of solar modulation ($\phi=550$ (solid curve) and 800 MV (dashed curve)) corresponding to the epoch when the experimental data were taken, and [b] relative elemental abundances compared with previous observations. The experimental data is compared with the results of a propagation model GALPROP. See text for details. (Symbols refer to CRIS: {\large $\bullet$}, Maehl et al.\1977:$\Box$, Webber et al.\1972:{\normalsize $\triangle$}, Englemann et al.\1990:$\blacksquare$, Orth et al.\1978:{\small $\blacklozenge$}, Buffington et al.\1978:{\small $\lozenge$}, Simon et al.\1980:\sidediamond, Buckley et al.\1994:{\small \ellipse}, Chapel and Webber 1981: {\small $\bigtriangledown$}, Lezniak et al.\1978:{\small $\triangle$}, Muller et al.\1991:$\dagger$, Dwyer et al.\1978:{\large $\triangleright$}, Webber et al.\1977:$\blacktriangleright$, Garcia-Munoz et al.\1987:$\blacktriangleleft$, Duvernois et al.\1996:$\ptimes$, Mewaldt et al.\1981:$\lowerdiamond$, Wiedenbeck and Greiner 1980: \crossa, Webber et al.\2002:$\blacktriangledown$, Hagen et al.\1977:$\bigstar$, Fisher et al.\1976:{\small $\upperbox$}, Juliusson et al.\1974:{\bf +}, Garcia-Munoz et al.\1977:{\large $\circ$}, Connell et al.\2001:{\normalsize $\times$}, Krombel et al.\1988:{\large $\triangleleft$}, Lukasiak et al.\1999:$\blacktriangle$).}
\caption{Number of muons (\lleft) and number of electrons (\rright) at shower maximum as function of energy for primary protons and iron nuclei according to CORSIKA simulations (symbols). The lines are predictions according to \eref{nmueq} and \eref{neeq}, respectively, for protons (\line) and iron (\dashed) nuclei.}
\caption{Average depth of the shower maximum for primary photons, protons, and iron nuclei according to CORSIKA simulations. The lines indicate predictions according to \eref{xmaxp} (\dashed) and the same function shifted up by 110~\gcm2 (\line).}
\caption{The X-ray to optical/nIR broad-band spectrum of GRB\,050716, 0.04 days post burst. The best fit is shown for a broken power~law connecting the two wavelength ranges, with host-galaxy absorption for an SMC-like extinction curve at$z = 2$. The X-ray spectrum has been represented relatively simply by a single data point. This has the purpose of enhancing the \chisqred\sensitivity on the nIR data instead of on the X-ray data, since we try to calculate the estimated optical/UV host-galaxy extinction.\label{fig:bbspec} }
\caption{({\em a}) The $4/3$ law for a run with $\Pm=1250$, $\Re\sim1$ (run S5 of \citealt{schek04}: $256^3$, $\nu=5\times10^{-2}$, $\eta=4\times10^{-5}$). We plot minus the left- (\pluses) and the right-hand (\solid) sides of \eqref{eq:ft}, and also minus the dissipative term (\dotted). Several individual third-order structure functions in the left-hand side of \eqref{eq:ftuB} are shown: $-\avg{\dzl^-|\dvz|^2}$ (\boldsolid), $2\avg{\dBl\dvB\cdot\dvu}$ (\dashdotted), $-\avg{\dul|\dvu|^2}$ (\dashed), $-\avg{\dul|\dvB|^2}$ (\dashtridot). The three functions that are not shown are small. ({\em b}) The second-order structure function $\avg{|\dvB|^2}/2\avg{B^2}$ for a sequence of runs with increasing $\Pm$ and constant $\Re\sim1$ (runs S1--S6 of \citealt{schek04}).}
\caption{ ({\em a}) The $4/5$ law for a run with $\Pm=1$, $\Re\simeq 400$ (run A of \citealt{schek04}: $256^3$, $\nu=\eta=5\times10^{-4}$). We plot minus the left- (\pluses) and the right-hand (\solid) sides of \eqref{eq:ff}, minus the viscous term (\dotted), $-\avg{\dul^3}$ (\dashed) and $6\avg{\Bl^2\dul}$ (\dashdotted). Inset: plot of $\lstar = -6\avg{\Bl^2\dul}r/\avg{\dul^3}$. ({\em b}) The $4/3$ law for the same run. See caption of \figref{fig:stokes}{a} for explanation of line styles. The functions $\avg{\dul\dvu\cdot\dvB}$, $\avg{\dBl|\dvu|^2}$ and $\avg{\dBl|\dvB|^2}$ are again small (not shown).}
\caption{Scale-by-scale budget for the isotropic part of Kolmogorov equation: log-log plots of $-\moy{U_R}^0_0$ (\longdashline), $2\alpha g/r^2\int_0^r y^2\moy{ \delta u_z\delta \theta}^0_0\mbox{d}y$ (\dashline), $2\nu\partial_r\moy{(\delta u_i)^2}^0_0$ (\dashdotline), $\moy{NH}^0_0$ (\dashtripledotline). The sum of the l.h.s. terms of equation~(\ref{eqVfinal_l=0}) (\fullline) matches the r.h.s. ``4/5 term'' (\dotline).\label{fig:budgets}}
\caption{Left: spherical harmonics spectrum of $\moy{U_R}$ at the centre of the box for $r/\eta=7.9$ (\fullline), $r/\eta=15.3$ (\dotline) and $r/\eta=22.6$ (\dashline). %Each spectrum is normalized % with respect to $\moy{U_R}^0_0(r)$. Right: local log-slope of $\moy{U_r}(r,\pi/2)$ (\fullline) and using spherical harmonics up to $\ell=0$ (+), $\ell=2$ ($\times$), $\ell=4$ ($\diamond$), $\ell=6$ ($\triangle$) (see equation~(\ref{eq:lincomb})). \label{fig:specandproj}}
\caption{\label{t-d-offset} The depth of the groove for $h=10$~nm as a function of time (log--log scale). {\color{green}$\bullet$}:~$r = 1$, {\color{gray}$\blacksquare$}:~$r = 3\times10^{-2}$, {\color{blue}$\Box$}:~$r = 10^{-2}$, {\color{brown}$\circ$}:~$r = 3\times10^{-3}$, and {\color{charcoal}$+$}:~$r = 10^{-3}$. The curves are offset by a half decade, for clarity. The lines are power law fits to Eq.~(\ref{eq-depth}) with parameters as shown in Table~\ref{table-alpha}.}
\caption{\label{t-cMax-slow3} $\cHat$ as a function of time for $r=10^{-3}$. {\color{teal}$+$}:~$h =5$~nm, {\color{blue}$\times$}:~$h =6$~nm, {\color{violet}$\Box$}:~$h =8$~nm, {\color{charcoal}$\triangle$}:~$h =10$~nm, {\color{green}$\blacksquare$}:~$h =12$~nm, {\color{red}$\blacktriangle$}:~$h =14$~nm, and {\color{brown}$\bullet$}:~$h =20$~nm.}
\caption{Propagation of an antikink with $m=-2$, $\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}$ and $\stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}$ within an $800\cdot401\cdot2$-point grid defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-40,40\right] \wedge y\epsilon\left[-40,40\right] \right\}$.} \label{fig3} \end{figure} \subsubsection{Propagation}\label{311} \par The dependence on the $m$ parameter of the kink solution is studied. Regardless of the value of $m$, a nonlinear trail dued to numerical reasons is always present in the propagation of the kink. As $m$ rises, the trail becomes more significant, which may be explained by the steepening of the step. Indeed, $\phi\propto arctg\left(e^{\frac{x}{\sqrt{1-\beta^2}}}\right)=arctg\left(e^{|m|x}\right)$ $\left(\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}, \stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}\right)$. For $m=7$ or higher, the kink is destabilised and no proper propagation is achieved. Another important $m$-dependent feature is the propagation velocity. The theoretical motion is given by $x=x_0+\beta t$ $\left(if \stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}\right)$; so, the velocity should be constant and equal to $\beta$. To determine this velocity numerically, several pairs of time and position of the center point are registered and a linear fit is applied using \textit{Origin 5.0} \textregistered - an example is presented in figure \ref{fig4}. The results for $m=\left\{1.5,2.0,2.5,3.0\right\}$ $\left(using \stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}, \stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}\right)$ are presented in table \ref{table1}. It is obvious that greater values of $m$ lead to less precision and exactness on the determination of the velocity. Taking into account the last column of table \ref{table1}, the most favourable situation occurs when $m=2.0$. Therefore, the next sections will use this kind of kinks (and the corresponding antikinks with $m=-2.0$) as they are less disturbed by numerical errors. Moreover, the fitted values of $x_0$ are close to 0, which means the law $x=x_0+\beta c t$ is being followed $\left(\stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}\right)$. \par It is interesting to note that all velocities in table \ref{table1} are below 1 ($c=1$), which is the typical velocity of a wave in the case $F(\phi)=0$. As $m\rightarrow \infty$, the velocity of the kink tends to 1 since $\beta\rightarrow 1$. %%%%FIGURE 4 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/linearfit} \caption{Linear fit of pairs $\left\{time,position\right\}$ of the center point of a kink with $m=2$, $\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}$ and $\stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}$ propagating within an $800\cdot401\cdot2$-point grid defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-40,40\right] \wedge y\epsilon\left[-40,40\right] \right\}$.} \label{fig4} \end{figure} %%%%TABLE 1 \begin{table}[h] \begin{center} \begin{tabular}{c|c|c|c|c} \hline\hline m [] & $x_0^{(num)}$ [] & $\beta ^{(num)}$ [] & $\beta ^{(teo)}$ [] & $\delta_{\beta}$ [\%] \\ \hline 1.5 & -0.093 $\pm$ 0.000 & 0.743 $\pm$ 0.000 & 0.745 & 0.28 \\ \hline 2.0 & -0.108 $\pm$ 0.000 & 0.864 $\pm$ 0.000 & 0.866 & 0.26 \\ \hline 2.5 & -0.081 $\pm$ 0.028 & 0.908 $\pm$ 0.001 & 0.917 & 0.97 \\ \hline 3.0 & 0.017 $\pm$ 0.039 & 0.922 $\pm$ 0.002 & 0.943 & 2.25 \\ \hline \hline \end{tabular}\caption{Determination of the propagation velocity of kinks with $m=\left\{1.5,2.0,2.5,3.0\right\}$, $\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}$ and $\stackrel{\rightarrow}{r_0}=\stackrel{\rightarrow}{0}$ propagating within an $800\cdot401\cdot2$-point grid defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-40,40\right] \wedge y\epsilon\left[-40,40\right] \right\}$. The last column shows the relative difference between $\beta^{(num)}$ and $\beta^{(teo)}$: $\delta_{\beta}=\frac{\left|\beta^{(num)}-\beta^{(teo)}\right|}{\beta^{(teo)}}\cdot 100$.}\label{table1} \end{center} \end{table} \subsubsection{Collision} \par The collision between kink-like solitons is achieved by superposing them. However, in general, this procedure is not legitimate since the sine-Gordon equation \eqref{eqSG} is nonlinear. The condition for the coexistence of a pair of legitimate kink-like solitons is \begin{equation}\label{superpose} sin(\phi_1+\phi_2)=sin\phi_1+sin\phi_2 \end{equation} where $\phi_1$ and $\phi_2$ are solutions of \eqref{eqSG}. Thus, two kink-like solitons may be superposed if their steps are sufficiently separated, because in that case on each grid point one of the two solutions is a multiple of $2\pi$, which means \eqref{superpose} is verified. In this way, the kink-kink collision, shown in figure \ref{fig5}, can be easily started and studied. Only the collision of kinks with symmetric velocity and same $|m|$ is simulated. Future works may take into account other situations. %%%%%FIGURE 5 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/kinkkink} \caption{Kink-kink collision. Both kinks have $m=2$ and $\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}$; the kink propagating left-to-right (K$\rightarrow$) has $\stackrel{\rightarrow}{r_0}=(-10,0)$, while the one propagating right-to-left (K$\leftarrow$) has $\stackrel{\rightarrow}{r_0}=(10,0)$. An $800\cdot401\cdot2$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-40,40\right]\wedge y\epsilon\left[-40,40\right]\right\}$.} \label{fig5} \end{figure} \par Figure \ref{fig5} shows that the kinks emerge intact (shape-wise) from collisions, which is typical of a soliton's behaviour. However, as they are nonlinear objects, they interact with each other during collisions. Evidence of this interaction is looked for in solitons velocity. The determination of the velocity follows the procedure explained in section \ref{311}, but picking pairs of time and position only after the interaction. The results obtained for each of the kinks are presented in table \ref{table2}. Comparing these results with those referring to propagation only (table \ref{table1}, line with $m=2.0$), one understands that the value of $\beta$ does not change significantly, while $x_0$ is now clearly above 0. This means that the interaction speeds up each one of the kinks, but does not alter their velocity afterwards. %%%%% TABLE 2 \begin{table}[h] \begin{center} \begin{tabular}{c|c|c} \hline\hline & $x_0^{(num)}$ [] & $\beta^{(num)}$ [] \\ \hline K$\rightarrow$ & 0.703 $\pm$ 0.000 & 0.864 $\pm$ 0.000 \\ \hline K$\leftarrow$ & 0.703 $\pm$ 0.000 & 0.864 $\pm$ 0.000 \\ \hline \hline \end{tabular} \caption{Determination of the propagation velocity of the kinks in the kink-kink collision referenced by figure \ref{fig5}.}\label{table2} \end{center} \end{table} \subsection{Light bullets} \label{32} \par The \textit{light bullet} \cite{Povich,Xin} represents a well-localised two-dimensional moving pulse. Its previewed space-time dependence is \begin{eqnarray}\label{lb} \phi\left(t,\stackrel{\rightarrow}{r}\right)=&sin& \left(\gamma \stackrel{\rightarrow}{d} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{r_0}\right)+\omega t\right)\cdot \nonumber \\ & \cdot e&^{-\frac{1}{4\sigma^2}\left(\left[\stackrel{\rightarrow}{d} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{r_0}\right)\right]^2+\left[\stackrel{\rightarrow}{u} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{r_0}\right)\right]^2\right)} \end{eqnarray} where $\omega=\sqrt{1+\gamma^2}$ and $\stackrel{\rightarrow}{u}$ is a normalised direction perpendicular to $\stackrel{\rightarrow}{d}$. \subsubsection{Propagation}\label{321} \par The evolution of a single light bullet is represented in figure \ref{fig11}. There is no significant modification of shape - light bullets behave indeed as solitons. Nevertheless, a nonlinear trail is formed and the $\sigma$-value of the envelope function changes during propagation. To quantify the former, one compares the integral of $\phi$ in the trail region to that in the pulse region: \begin{displaymath} \epsilon=\frac{\int_{trail} \phi dx dy}{\int_{pulse} \phi dx dy} \end{displaymath} The higher $\epsilon$, the more important the trail is in comparison with the soliton. It is important to note that the trail is dued not only to numerical reasons but also to the fact that \eqref{lb} is not an analytical solution of \eqref{eqSG}, unlike the case of the kink-like solitons. The latter effect is evaluated fitting the numerical values of $\phi$ after propagation to \begin{displaymath} a_0+a_1e^{-\frac{1}{4 {a_2}^2}\left(\left[\stackrel{\rightarrow}{d} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{a_3}\right)\right]^2+\left[\stackrel{\rightarrow}{u} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{a_3}\right)\right]^2\right)} \end{displaymath} where sometimes not all parameters are free to vary for convergence purposes. %%%% FIGURE 11 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/lbprop} \caption{Propagation of a light bullet with $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_0}=(-15,0)$, $\stackrel{\rightarrow}{d}=\stackrel{\rightarrow}{e_x}$ and $\stackrel{\rightarrow}{u}=\stackrel{\rightarrow}{e_y}$. An $800\cdot151\cdot15$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-30,30\right]\right\}$. The images are the projection of the solution in the xz plane. In this case, $\Gamma=7.85$.} \label{fig11} \end{figure} \par One can now study $\epsilon$ and $a_2$ as functions of $\gamma$ and $\sigma$ and then choose the most favourable pair $\left\{\gamma,\sigma\right\}$ to perform collisions between light bullets. Tables \ref{table4} and \ref{table5} present the results of these simulations. Although the case $\left\{\gamma,\sigma\right\}=\left\{1.0,5.0\right\}$ is the one with less significant trail, it leads to a strong deviation of $\sigma$ after propagation. The pair of parameters $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$ seems to be the one in which less numerical errors occur; therefore, these are the values used in the next sections. %%%%TABLE 4 \begin{table}[!h] \begin{center} \begin{tabular}{c|c|c|c} \hline\hline $\sigma$ [] $\backslash$ $\gamma$ [] & 1.0 & 2.0 & 3.0 \\ \hline 1.0 & 0.0402 & 0.0439 & 0.0447\\ \hline 2.5 & 0.0448 & 0.0430 & 0.0442\\ \hline 5.0 & 0.0380 & 0.0487 & 0.0461\\ \hline $\Gamma$ [] & 15.71 & 7.85 & 5.23 \\ \hline\hline \end{tabular}\caption{Values of $\epsilon$ for light bullets with $\gamma=\left\{1.0,2.0,3.0\right\}$ and $\sigma=\left\{1.0,2.5,5.0\right\}$ propagating within an $800\cdot151\cdot15$-point grid defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-30,30\right]\right\}$. $\Gamma$-values are also shown for each $\gamma$ used.}\label{table4} \end{displaymath}\end{center} \end{table} %%%%TABLE 5 \begin{table}[!h] \begin{center} \begin{tabular}{c|c|c|c} \hline\hline $\sigma$ [] $\backslash$ $\gamma$ [] & 1.0 & 2.0 & 3.0 \\ \hline 1.0 & 2.688 & 1.156 & 2.067 \\ \hline 2.5 & 2.704 & 2.468 & 2.573 \\ \hline 5.0 & 5.907 & 4.596 & 4.639 \\ \hline $\Gamma$ [] & 15.71 & 7.85 & 5.23 \\ \hline\hline \end{tabular}\caption{Values of the fitting parameter $a_2$ for light bullets with $\gamma=\left\{1.0,2.0,3.0\right\}$ and $\sigma=\left\{1.0,2.5,5.0\right\}$ propagating within an $800\cdot151\cdot15$-point grid defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-30,30\right] \wedge y\epsilon\left[-30,30\right] \right\}$. $\Gamma$-values are also shown for each $\gamma$ used.}\label{table5} \end{center} \end{table} \subsubsection{Collision} \par Only the collision between light bullets of $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$ is simulated. Other cases may be delt with in future approaches. The numerical method is started by superposing light bullets according to \eqref{superpose}. Figure \ref{fig12} shows a head-on collision. The solitons interact with each other and emerge essentially intact and with no change in propagation direction. However, the $\sigma$-value of the envelope function of each bullet does not behave as in the propagation case. In fact, the value of the fitting parameter $a_2$ after collision is 2.09. It seems that head-on collisions lead to a shrink of the soliton, but this conclusion must be tested by further and more complete simulations. %%%%FIGURE 12 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.4]{images/180collision} \caption{Head-on collision between two light bullets with $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(-15,0)$ and $\stackrel{\rightarrow}{r_{02}}=(15,0)$. An $800\cdot151\cdot15$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-30,30\right]\right\}$. The images are the projection of the solution in the xz plane.} \label{fig12} \end{figure} \subsection{Collision between light bullets and other solutions}\label{33} \par In order to study the interaction of light bullets with kink-like solutions, kink-, antikink- and standing kink-light bullet head-on collisions are set up and shown, respectively, in figures \ref{fig13}, \ref{fig14} and \ref{fig15}. In all three cases, the light bullet seems to emerge intact from the interaction and continues its motion with no visible change. The velocities of the light bullet in the propagation case (section \ref{321}) and in kink-light bullet head-on collision are roughly determined and no significant difference is noticed. \par Moreover, a 30$^{\circ}$-collision between a light bullet and a standing kink is simulated - see figure \ref{figLBSK}. As in the previous scenarios, no modification in the light bullet propagation direction is detected, which evidences the robustness of these objects. %%%%FIGURE 13 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/k-lb} \caption{Kink-light bullet head-on collision. The light bullet has $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(-15,0)$ and $\stackrel{\rightarrow}{d_1}=\stackrel{\rightarrow}{e_x}$, while the kink presents $m=2$, $\stackrel{\rightarrow}{r_{02}}=\stackrel{\rightarrow}{0}$ and $\stackrel{\rightarrow}{d_2}=-\stackrel{\rightarrow}{e_x}$. An $800\cdot251\cdot31$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-35,35\right]\right\}$ and this domain is zoomed into $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-20,15\right] \wedge y\epsilon\left[-15,15\right] \right\}$. In the sequence of shots time flows from left to right and downwards. In this case, $\Gamma=13.09$.} \label{fig13} \end{figure} %%%%FIGURE 14 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/ak-lb} \caption{Antikink-light bullet head-on collision. The light bullet has $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(15,0)$ and $\stackrel{\rightarrow}{d_1}=-\stackrel{\rightarrow}{e_x}$, while the antikink presents $m=-2$, $\stackrel{\rightarrow}{r_{02}}=\stackrel{\rightarrow}{0}$ and $\stackrel{\rightarrow}{d_2}=-\stackrel{\rightarrow}{e_x}$. An $800\cdot251\cdot31$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-30,30\right] \wedge y\epsilon\left[-35,35\right] \right\}$ and this domain is zoomed into $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-15,20\right]\wedge y\epsilon\left[-15,15\right]\right\}$. In this case, $\Gamma=13.09$.} \label{fig14} \end{figure} %%%%FIGURE 15 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/sk-lb} \caption{Standing kink-light bullet head-on collision. The light bullet has $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(-15,0)$ and $\stackrel{\rightarrow}{d_1}=\stackrel{\rightarrow}{e_x}$, while the standing kink ($m=1$) presents $\stackrel{\rightarrow}{r_{02}}=\stackrel{\rightarrow}{0}$ and $\stackrel{\rightarrow}{d_2}=\stackrel{\rightarrow}{e_x}$. An $800\cdot251\cdot31$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-35,35\right]\right\}$ and this domain is zoomed into $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-20,15\right] \wedge y\epsilon\left[-15,15\right] \right\}$. In this case, $\Gamma=13.09$.} \label{fig15} \end{figure} %%%%FIGURE \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/sk-lb30} \caption{Standing kink-light bullet 30$^{\circ}$-collision. The light bullet has $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(-15cos30^{\circ},-15sin30^{\circ})$ and $\stackrel{\rightarrow}{d_1}=cos30^{\circ}\stackrel{\rightarrow}{e_x}+sin30^{\circ}\stackrel{\rightarrow}{e_y}$, while the standing kink ($m=1$) presents $\stackrel{\rightarrow}{r_{02}}=\stackrel{\rightarrow}{0}$ and $\stackrel{\rightarrow}{d_2}=\stackrel{\rightarrow}{e_x}$. An $800\cdot61\cdot61$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-30,30\right] \wedge y\epsilon\left[-30,30\right] \right\}$ and this domain is zoomed into $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-20,20\right]\wedge y\epsilon\left[-20,20\right]\right\}$. In this case, $\Gamma=3.14$.} \label{figLBSK} \end{figure} \par Another collision taken into account is the one between a light bullet and a \textit{standing breather}, shown in figure \ref{fig16}. The latter object is an analytical solution of \eqref{eqSG} which is oscillatory and may be interpreted as a bound state between a kink and an antikink. Its behaviour is previewed analytically in the (1+1)-dimension case \cite{Infeld} and can be extended to \begin{equation}\label{stdbreather} \phi\left(t,\stackrel{\rightarrow}{r}\right)=4 arctg\left(\frac{m}{\sqrt{1-m^2}}\frac{sin\omega t}{cosh\left(m\stackrel{\rightarrow}{d} \cdot \left(\stackrel{\rightarrow}{r}-\stackrel{\rightarrow}{r_0}\right)\right)}\right)\end{equation} where $\omega=\frac{2\pi}{T}=\sqrt{1-m^2}$ and $|m|<1$. In this collision, the light bullet still emerges intact, although the standing breather is completely ruined after the interaction. In other words, the bound state kink-antikink is destroyed by the light bullet, even though this does not destroy isolated kinks nor antikinks. %%%%FIGURE 16 \begin{figure}[h]%CUIDADO COM ESTE h \centering %centra figura \includegraphics[scale=0.43]{images/sb-lb} \caption{Standing breather-light bullet head-on collision. The light bullet has $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$, $\stackrel{\rightarrow}{r_{01}}=(-10,0)$ and $\stackrel{\rightarrow}{d_1}=\stackrel{\rightarrow}{e_x}$, while the standing breather presents $m=0.8$, $\stackrel{\rightarrow}{r_{02}}=\stackrel{\rightarrow}{0}$ and $\stackrel{\rightarrow}{d_2}=\stackrel{\rightarrow}{e_x}$. An $800\cdot251\cdot31$-point grid is defined in $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right]\wedge x\epsilon\left[-30,30\right]\wedge y\epsilon\left[-35,35\right]\right\}$ and this domain is zoomed into $\left\{\left(t,x,y\right): t\epsilon\left[0,100\right] \wedge x\epsilon\left[-25,25\right] \wedge y\epsilon\left[-15,15\right] \right\}$. In this case, $\Gamma=13.09$.} \label{fig16} \end{figure} \par Figures \ref{fig13}, \ref{fig14}, \ref{fig15}, \ref{figLBSK} and \ref{fig16} all together allow one to confirm the robustness of light bullets observed in \ref{32}. Indeed, these solutions pass through analytical solutions \eqref{1sol} and \eqref{stdbreather} and remain unchanged. This property may have interesting consequences, specially if one identifies light bullets with optical pulses propagating in different media \cite{Xin}. \section{Final remarks}\label{4} \par The analysis of kink and antikink solitons revealed an obvious shape consistency during propagation, which is typical of solitons behaviour. Moreover, an $m$-dependent study of the velocity in the kink case was carried out and it followed the theoretical solution \eqref{1sol}, as expected. \par As for kink-kink collision, one verified that kinks do not alter their shape, propagation direction nor velocity after head-on collisions with each other. Nevertheless, the numerical value of $x_0$ was found to be significantly above 0 for both kinks, which proves that nonlinear interaction has taken place and that it momentarily speeded up both solitons. \par In the light bullet case, the propagation was characterised as a function of the parameters $\gamma$ and $\sigma$ and the most favourable case was identified: $\left\{\gamma,\sigma\right\}=\left\{2.0,2.5\right\}$. With this configuration, a head-on collision was set up and nonlinear effects in the shape of the light bullets were measured (through $a_2$). The shape consistency of the light bullets noticed in the latter collision was also seen in head-on collisions with kink, antikink, standing kink and standing breather. \par Moreover, a 30$^{\circ}$-collision between a light bullet and a standing kink was set up as well and it reinforced the idea that light bullets and kinks are indeed robust objects. We also found that the direction of the propagation of the light bullet is maintained after it collides with the kink. \par Future developments may study the dependence on $\gamma$ and $\sigma$ of light bullet collisions and verify if these lead to a shrink of the light bullets or not. Another interesting point is to understand why kink, antikink and standing kink survive to the collision with a light bullet while the standing breather does not. % The Appendices part is started with the command \appendix; % appendix sections are then done as normal sections % \appendix % \section{} % \label{} %\begin{thebibliography}{00} % \bibitem{label} % Text of bibliographic item % notes: % \bibitem{label} \note % subbibitems: % \begin{subbibitems}{label} % \bibitem{label1} % \bibitem{label2} % If there is a note, it should come last: % \bibitem{label3} \note % \end{subbibitems} %\bibitem{} %\end{thebibliography} \begin{thebibliography}{99} \bibitem{SGS} \textit{Solitons in the sine-Gordon equation}; Nonlinear Science; Los Alamos Science Special Issue 1987 \bibitem{DiGarbo} Di Garbo, A. et al; \textit{Dynamical Properties of a Kink of the sine-Gordon Equation Trapped in a Potential Well}; Nonlinear Analysis 47 (2001) 5967-5978 \bibitem{pion} http://valdostamuseum.org/hamsmith/SnGdnPion.html \bibitem{Sobolev} Sobolev, A., Pankratov, A. and Mygind, J.; \textit{Numerical simulation of the self-pumped Long Josephson junction using a modified Sine-Gordon model}; preprint submitted to Elsevier Science; 10th September 2005 \bibitem{Wallraff} Wallraff, A., Ustinov, A., Kurin, V., Shereshevsky I. and Vdovicheva, N.; \textit{Whispering Vortices}; Physical Review Letters; volume 84; number 1; 3rd January 2000 \bibitem{artgall} http://homepages.tversu.ru/\~{}s000154/collision/main.html \bibitem{Povich} Povich, T. and Xin, J.; \textit{A Numerical Study of the Light Bullet Interaction in the (2+1) Sine-Gordon Equation}; Journal of Nonlinear Science; volume 15; 11-25; 2005 \bibitem{Xin} Xin, J.; \textit{Modeling light bullets with the two-dimensional sine-Gordon equation}; Physica D 135 (2000) 345-368 \bibitem{Infeld} Infeld, E. and Rowlands, G.; \textit{Nonlinear waves, solitons and chaos}; Cambridge University Press; 1990 \bibitem{Whitham} Whitham, G.B.; \textit{Linear and Nonlinear Waves}; John Wiley \& Sons Inc; 1974\end{thebibliography} \end{document} }\end{figure}}\end{equation}}\end{center}\end{table}}\end{eqnarray}}
\begin{caption}{Solution branch for 3-fold rotationally-symmetric travelling waves plotted on a friction factor versus axial wavenumber plot at $Re=2400$. The blue dotted line represents the lower bound given by the Hagen-Poiseuille solution ($\Lambda=64/Re$) and the upper blue dashed line corresponds to the $Re=2400$ value of the log-law parametrisation of experimental data $\frac{1}{\sqrt{\Lambda}}=2.0 \log (Re_m \sqrt{\Lambda})-0.8$ (see Schlichting 1968 equation (20.30)). The solution branch is only shown as far as it is assured resolved (hence the loose ends: the main mapping resolution was $(8,30,6)$ in the truncation nomenclature of Wedin \& Kerswell 2004). The dotted vertical lines indicate the wavenumbers ($\alpha=0.625n$ in units of $2/D$, $n=1,2,3,4,5$) which fit into a pipe of length $\pi/0.625\, D$ long. The letters are used to label each allowable TW together with the wavenumber. \label{fig:3}} \end{caption}
\begin{caption}{ The surplus kinetic energy per unit mass, in units of $U^2$ versus wall shear stress $\tau$ in units of $2 \rho U^2/Re$ for $\bu_{DNS}$ starting at TW $4b\_3.125$ (marked as a {\color{black}$\circ$} with a $\cross$). The solid line indicates the turbulent evolution for one sign of the eigenvalue perturbation and the thick dashed line traces out the uneventful relaminarisation for the other (respectively $4b\_3.125(+/-)$). The laminar state is represented by the point $(-4,0)$. All the TWs present are also plotted: {\color{blue}$\triangle$} for 2-fold TWs, {\color{red}$+$} for 3-fold TWs and {\color{black}$\circ$} for 4-fold TWs. Filled in symbols indicate TWs which appear to be visited by $\bu_{DNS}$ and the numbered dots indicate the temporal points of closest approach. In chronological order: 1 - $4f\_3.125$ (lower black filled circle); 2 - $2b\_0.625$ (blue filled triangle {\em furthest} from 3); 3 - $2a\_1.25$ (closer blue filled triangle); 4 - $4c\_3.125$ (upper black filled circle) and 5 - $3a\_3.125$ (circled red $+$). } \label{ketau_4b} \end{caption}
\begin{caption}{ The surplus kinetic energy per unit mass in units of $U^2$ versus wall shear stress $\tau$ in units of $2 \rho U^2/Re$ for $\bu_{DNS}$ starting at the upper branch TW $3b\_3.125(+)$ (marked as a {\color{red}$\square$} with a {\color{red}$+$} in it). The laminar state is represented by the point $(-4,0)$. All the TWs present are also plotted: {\color{blue}$\triangle$} for 2-fold TWs, {\color{red}$+$} for 3-fold TWs and {\color{black}$\circ$} for 4-fold TWs. The numbered dots indicate the temporal points of closest approach. In chronological order: 1 - $3b\_2.5$ (rightmost red circled $+$); 2 - $3j\_2.5$ (leftmost red circled $+$); 3 - $3c\_2.5$ (middle red circled $+$); 4 and 6 - $2a\_1.25$ (large blue triangle); 5 and 7 - $2b\_1.875$ (blue solid triangle). } \label{ketau_3b} \end{caption}
\caption{Shear rate dependence of the viscosities of the fluids: ({\textcolor{blue}{$\circ$}})~displaced fluid~2 with $pH \approx 3$, ({\textcolor{black}{$\Box$}})~displacing Fluid 1 , ({\textcolor{red}{$\triangle$}})~displaced Fluid 2 with $pH \approx 7$. }
\caption{(a) Normalized width of the tip versus the normalized displacement distance, $\hat U_0 \hat t/\hat R$, for several values of the flow rate: ($\blacksquare$) $\hat {Q}=0.063$~ml/s, ({\textcolor{red}{$\bullet$}}) $\hat {Q}=0.145$~ml/s, ({\textcolor{blue}{$\blacktriangle$}}) $\hat {Q}=0.19$~ml/s, ({\textcolor{green}{$\blacktriangleleft$}}) $\hat {Q}=0.3$~ml/s. The experiments belong to Control sequence~$1$ (see table \ref{experiments}). (b) Normalized width of the tip versus the normalized displacement distance, $\hat U_0 \hat t/\hat R$, for several values of the flow rate: ({\textcolor{green}{$\blacktriangleleft$}}) $\hat {Q}=0.13$~ml/s, ($\blacksquare$) $\hat {Q}=0.18$~ml/s, ({\textcolor{brown}{$\blacktriangledown$}}) $\hat {Q}=0.2$~ml/s, ({\textcolor{red}{$\bullet$}}) $\hat {Q}=0.31$~ml/s, ({\textcolor{blue}{$\blacktriangleright$}}) $\hat {Q}=0.47$~ml/s. The experiments belong to the Reactive sequence (see Table~\ref{experiments}).}
\caption{Example results: $\sigma_R$ vs $\alpha$ for the base flow of Fig.~\ref{f.10} with the viscosity amplitudes ({\textcolor{blue}{$\blacktriangle$}})~$a=100$ ($\blacksquare$)~$a=5$ and ({\textcolor{red}{$\bullet$}})~$a=20$: (a) $\Sc = 10^4$. In the inset we display the dependence of the critical wave number, $\alpha_c$, on the amplitude $a$; (b) $\Sc = 10^6$.}
\caption{\label{fig:wt} %\label{fig:epsart} \footnotesize{\bf Influence of temperature and transient antisense interactions on co-transcriptional folding.} Equilibrium and native structures of reverse switch (R) with {\it in vitro} T7 transcription at 25$^\circ$C (left, see text) and under {\it in vitro} T7 transcription in presence of 0.3~nmol/$\mu$l of the 7-nt antisense DNA oligonucleotide CCTCTAC (right, see text). Structures are separated on a 12\% 19:1 acryl-bisacrylamide \textcolor[rgb]{0,0,0}{nondenaturing} gel (temperature $<$10$^\circ$C) and observed using ethidium bromide staining as on Fig.~2 (see Materials and Methods). }
\caption{\label{fig:wt} %\label{fig:epsart} {\footnotesize{\bf Antisense regulation of co-transcriptional folding paths.} Interpretation of the encoded (left) and redirected (right) co-transcriptional folding paths of the reverse switch (Fig.~4). \textcolor[rgb]{0,0,0}{This is based on simulations performed using the kinefold server\protect\cite{kinefold} ({\tt \footnotesize http://kinefold.curie.fr}); To simulate the effect of antisense interaction, the 7mer and RNA switch sequences are actually attached together via an inert linker (made of 'X' bases that do not pair). } }}
\caption{\label{fig-levels_mp_mf-1} (Color online) Many particle energies $E_n$ and mean-field eigenenergies ${\cal H}_{stat}$ (\textcolor{red}{- -}) as a function of the onsite energy $\varepsilon$ in the subcritical region ($g=-0.5/N_s$) for $v=1$ and $N=10$ particles ($\hbar=1$). The spectrum is organized by the mean-field eigenenergies ${\cal H}_{stat}$ (\textcolor{red}{- -}). The exact quantum eigenvalues shown for $\varepsilon\leq 0$ (\textcolor{green}{---}) are almost exactly reproduced by the semiclassical ones, $E_n^{sc}$, shown for $\varepsilon\geq 0$ (\textcolor{blue}{---}).}
\caption{\label{fig-levels_mp_mf-2} (Color online) Many particle energies $E_n$ and mean-field eigenenergies ${\cal H}_{stat}$ (\textcolor{red}{- -}) as a function of the onsite energy $\varepsilon$ in the supercritical region ($g=-3/N_s$) for $v=1$ and $N=10$ particles ($\hbar=1$). The spectrum is organized by the mean-field eigenenergies ${\cal H}_{stat}$ (\textcolor{red}{- -}). The exact quantum eigenvalues shown $\varepsilon\leq 0$ (\textcolor{green}{---}) are almost exactly reproduced by the semiclassical ones, $E_n^{sc}$, shown for $\varepsilon\geq 0$ (\textcolor{blue}{---}).}
\caption{\label{fig-Upm} (Color online) The potentials $U_-(p)$ (\textcolor{blue}{---}) and $U_+(p)$ (\textcolor{red}{- -}), for $N=10$ particles, $v=1$ and $\varepsilon=0.6$ in the subcritical ($g=-0.6/N_s$, top) and supercritical ($g=-4/N_s$, bottom) region, with $\hbar=1$.}
\caption{\label{fig-levels_mp_mf-N2} (Color online) Exact (- -), $E_n$ and semiclassical (\textcolor{red}{---}) many particle energies, $E_n^{sc}$, as a function of the onsite energy $\varepsilon$ in the subcritical region ($g=-0.9/N_s$) for $v=1$ and $N=2$ particles ($\hbar=1$).}
\caption{(color online) Time-averaged momentum distributions when quenching (a) from $V_0 = 0.5$ to $V=2$ %NEW (quantum critical point) and (b) from $V_0 = 1.5$ to $V=5$ (insulator) for $L=50$ sites. The time averages of two independent initial states with the same energy are compared to each other and to the thermal expectation value. In the right inset, results for $L=50$ (\textcolor{red}{$\ast$}) are compared to $L=100$ (\textcolor{red}{$\bigtriangledown$}) for the regions with the largest differences. As a reference, finite $T$ data for $L=50$ (dotted line) and $L=100$ (dashed line) are shown. Left insets: $\langle n_\pi \rangle$ vs. time t; the horizontal line is the finite $T$ value.}
\caption{Contour plot of $\log\left[\alpha_{xy}^{2d}/(2ek_B/h)\right]$. For small $T$ and $H$, an inter-layer Josephson coupling $J_\perp$, absent in these simulations, stabilizes 3d superconductivity. Note that, unlike the Nernst signal $e_N=\alpha_{xy}/\sigma_{xx}$ in Ref.~\cite{ong2}, there is no ridge field in our simulations of $\alpha_{xy}$. The difference is likely due to the diverging electrical conductivity $\sigma_{xx}$ as $T_c$ is approached. \label{fig:Acont}} \end{figure} {\em High Temperature Expansion:} For $T\ll J$, the phase-only model allows for an analytically tractable regime that is entirely different from the Gaussian regime considered previously\cite{Ussishkin1}. The high temperature expansion, carried out in powers of $J/k_BT$, is conveniently performed using the Martin-Siggia-Rose formalism\cite{MartinSiggiaRose,RaghuPodolskyVishwanath}. Since both $M_z$ and $\alpha_{xy}$ require a magnetic field, the expansion of these quantities involves graphs enclosing finite a magnetic flux. The leading term thus depends on the smallest closed graph -- on a square lattice this involves 4 links, and is hence proportional to $(J/T)^4$, whereas on a triangular lattice it goes as $(J/T)^3$: \begin{eqnarray} \alpha^{2d}_{xy}&=&\lambda\frac{2ek_B}{h}\left(\frac{J}{T}\right)^\mu\sin\frac{H}{H_0}\label{eq:largeT}\\ \frac{M_z^{2d}}{T}&=&-2\lambda\frac{2ek_B}{h}\left(\frac{J}{T}\right)^\mu\sin\frac{H}{H_0}\nonumber\\ \frac{|M_z^{2d}|}{T\alpha_{xy}^{2d}}&=&2.\label{eq:MoA} \end{eqnarray} Here, $\mu=4$ and $\lambda=\pi/8$ ($\mu=3$ and $\lambda=\pi/4$) for a square (triangular) lattice. Despite the lattice-dependent behavior of $\alpha_{xy}$ and $M_{z}/T$, their ratio (\ref{eq:MoA}) is equal to -2, independent of the lattice. The same value of -2 is obtained in the Gaussian regime of the dynamical Ginzburg-Landau equation\cite{Ussishkin2,Ussishkin1}, and seems to be a robust feature of fluctuating superconductivity at high temperatures. The ratio $|M_z|/T\alpha_{xy}$, shown in Fig.~\ref{fig:MoA}, is only weakly field-dependent, and tends to 2 at high-temperatures. This points to a close quantitative connection between the Nernst effect and diamagnetism \cite{ong1}. \begin{figure} \includegraphics[width=9cm]{MoA} \caption{The dimensionless ratio $|M_z|/T\alpha_{xy}$, as a function of magnetic field for various temperatures. In the high-temperature limit, this dimensionless ratio is expected to saturate at a value of 2 for all magnetic fields. {\it Inset:} $|M_z|/T\alpha_{xy}$ for $H=0.31 H_{0}$ vs. $T/T_{KT}$\label{fig:MoA}} \end{figure} {\em Comparison to Nernst Measurements in the Cuprates:} Experimental measurements of both Nernst voltage and conductivity are required to obtain $\alpha_{xy}$. Such experimental data is available on underdoped La$_{2-x}$Sr$_x$CuO$_4$ ($x=0.12$ and $T_c=28$ K) in weak fields. This is shown in Fig.~\ref{fig:yayu}, which displays the Nernst coefficient times conductivity, $\nu\sigma_{xx}=\left.\frac{d\alpha_{xy}}{dH}\right|_{H=0}$. To compare, we choose in our simulation, $J=J_l\equiv30.2$ K, corresponding to $T_c=1.04 T_{KT}$, and $H_0=50$ T (for this sample, $H_{c2}\approx$100 T, hence $H_{c2}>H_0$, consistent with a relatively dilute vortex liquid). With these values, the simulation gives good agreement with absolute experimental values in the regime $T_c<T<2T_c$ K, except for the very lowest temperature point $T=30$ K. This is very close to $T_c$, so that 3d superconducting fluctuations, ignored here, are likely dominant. {\em Comparison with High Temperature Data:} The inset of Fig.~\ref{fig:yayu} shows the measured $\nu\sigma_{xx}$ on a log-log plot extending to $T=120 K\approx 4T_c$. The data displays a rapid decay over a large temperature range, in general agreement with our expectations. In particular, a $T^{-4}$ decay is observed, which is the high temperature result (\ref{eq:largeT}) on a square lattice. However, since in the high temperature regime the precise power depends strongly on lattice geometry (in contrast to the intermediate temperature regime) justification for using a nearest neighbor square lattice model (as opposed to, say, a triangular lattice model) is required. While the underlying $d$-wave symmetry of the cuprates, coupled with the fact that $a$ is a microscopic length, about 6 times the lattice spacing, may be invoked, whether these are sufficient to justify the square lattice model is unclear at present and requires further work. The characteristic scale of $\alpha_{xy}$ at these temperatures is enhanced from what one would expect for a superconductor with $T_c=28K$. For example, the best fit to the square lattice high temperature expansion (solid line) requires $J=J_h=52 K$, larger than the effective coupling $J_l=30.2$ K that yiels the correct $T_c$. This may be naturally attributed to thermal d-wave quasiparticles, omitted in this analysis, which suppress the long distance superfluid density\cite{LeeWen} but not the superfluid density at shorter scales\cite{RaghuPodolskyVishwanath}, which controls the high temperature behavior. The ratio $J_h/J_l=1.6$ is consistent with measurements of the temperature dependent superfluid density in other cuprates \cite{lemberger}. A prediction from this scenario is that magnetization should continue to track $\alpha_{xy}T$. \begin{figure} \includegraphics[width=9cm]{ExpComparison} \caption{Comparison to underdoped La$_{2-x}$Sr$_x$CuO$_4$ ($x=0.12$, $T_c=28$ K). The experimental data\cite{ong2}, shifted by a constant quasiparticle background contribution $\nu_B\sigma_{xx}=-0.011 V/K\Omega Tm$, is indicated by $\bullet$. Simulation results are shown by X with error bars. The value of the experimental point at $T=30 K$ ($\blacksquare$) has been divided by 4 to fit in the figure. This point is very close to $T_c$, and hence $\nu\sigma_{xx}$ is likely dominated by 3d fluctuations, not considered in our simulations. \label{fig:yayu}} \end{figure} {\em Onset Temperature:} Ong and collaborators\cite{ong2} define a temperature $T_{\rm onset}$ where the fluctuating contribution to the Nernst effect can no longer be experimentally distinguished from the quasiparticle background. Here, since we have a natural scale for $\alpha_{xy}$, we define $T_{\rm onset}$ as the temperature where $\alpha_{xy}^{2d}$ has decayed to a small fraction $\delta$ of the quantum of thermoelectric conductivity, \begin{eqnarray} \alpha_{xy}^{2d}(T_{onset},H)=\frac{2ek_B}{h}\,\delta \end{eqnarray} For our model, $\alpha_{xy}^{2d}=\frac{2ek_B}{h}F\left(T/T_{KT},H/H_0\right)$, hence $T_{\rm onset}$ is proportional to $T_c$. The essential point is that, because $\alpha_{xy}^{2d}$ depends strongly on temperature, when inverted, $T_{onset}$ is only a weak function of $\delta$ and $H$. For instance, the choice $H=H_0/4$ and $\delta=0.01$ on the square lattice yields \begin{eqnarray} T_{\rm onset}\approx 3 T_c \end{eqnarray} This is consistent with the observation that the experimentally-defined $T_{\rm onset}$ roughly tracks $T_c$ as doping is varied\cite{ong2}. Due to the strong temperature dependence of $\alpha_{xy}$, the onset of Nernst effect is very sharp, in contrast to Gaussian fluctuations\cite{Ussishkin1}, where $\alpha_{xy}$ only decays as $1/(T_c-T)$ at high temperatures. Experimentally, measurements of the electrical conductivity $\sigma_{xx}$ do not have discernable contributions due to fluctuating superconductivity at temperatures of order $T_{\rm onset}$\cite{Orenstein}. Since $\sigma_{xx}$ is proportional to the parameter $\tau$ appearing in Eq.~(\ref{eq:Lang}), this places a constraint on the maximum value of $\tau$. The high temperature expansion yields \begin{eqnarray} \sigma^{2d}_{xx,\rm fluct}=\frac{4e^2}{h}\frac{\pi J_{\rm eff}^2}{4T^2}\frac{\tau}{\hbar}.\nn \end{eqnarray} As a benchmark, we note that BCS theory predicts the value $\tau_{\rm BCS}\approx 0.7 \hbar$. This yields a fluctuating conductivity at $T=2T_c$ that is only about $10\%$ of the quasiparticle conductivity in this material. In conclusion, we have studied the transverse thermoelectric conductivity $\alpha_{xy}$ and the diamagnetic response $M^z$ in the classical XY model with model-A dynamics. We have obtained numerical results at low temperatures, and analytic results at high temperatures, that are functions only of two variables, $T/T_c$ and $H/H_0$, where $H_0<H_{c2}$ is a characteristic field scale set by vortex parameters. In our model, we predict that $\alpha_{xy}$ and $M^z$ for different systems (e.g. different dopings) should collapse into a single curve when expressed in terms of the system-dependent $T_c$ and $H_0$. We show that $M_z/T$ and $\alpha_{xy}$ track each other and, in particular, we predict that their ratio tends to $-2$ at high temperatures. Measurements of $\alpha_{xy}$ on the underdoped cuprate La$_{1.88}$Sr$_{0.12}$CuO$_4$ display a sharp temperature decay, in agreement with our model. We would like to thank P.W. Anderson, D. Huse, S. Kivelson, J.E. Moore, S. Mukerjee, N.P. Ong, and Y. Wang for many insightful discussions. \begin{thebibliography}{2} \bibitem{Orenstein} J. Corson, R. Mallozzi, J. Orenstein, J. N. Eckstein, and I. Bozovic, Nature {\bf 398}, 221 (1999) \bibitem{ong1} Y. Wang, {\it et al.}, Phys. Rev. Lett. {\bf 95}, 247002 (2005) \bibitem{ong2} Y. Wang, L. Li, and N.P. Ong, Phys. Rev. B {\bf 73}, 024510 (2006) \bibitem{behnia1} R. Bel, K. Behnia, and H. Berger, Phys. Rev. Lett. {\bf 91}, 066602 (2003) \bibitem{behnia2} A. Pourret {\it et al.}, Nature Physics {\bf 2}, 683 (2006) \bibitem{sondheimer} E.H. Sondheimer, Proc. Roy. Soc. A {\bf 193}, 484 (1948) \bibitem{EmeryKivelson} V.J. Emery, and S.A. Kivelson, Nature {\bf 374}, 434 (1995) \bibitem{FisherFisherHuse} D.S. Fisher, M.P. A. Fisher, and D. A. Huse, Phys. Rev. B {\bf 43}, 130 (1991) \bibitem{anderson} P.W. Anderson, Nature Physics (2006) \bibitem{Ussishkin1} I. Ussishkin, S. L. Sondhi, and D. A. Huse, Phys. Rev. Lett. {\bf 89}, 287001 (2002) \bibitem{Ussishkin2} I. Ussishkin, unpublished notes \bibitem{Mukerjee1} S. Mukerjee and D. A. Huse, Phys. Rev. B {\bf 70}, 014506 (2004) \bibitem{CooperHalperinRuzin} N. R. Cooper, B. I. Halperin, and I. M. Ruzin, Phys. Rev. B {\bf 55}, 2344 (1997) \bibitem{HohenbergHalperin} P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. {\bf 49}, 435 (1977) \bibitem{OganesyanHuseSondhi} V. Oganesyan, D.A. Huse, and S. L. Sondhi, Phys. Rev. B {\bf 73}, 094503 (2006) \bibitem{MartinSiggiaRose} P. C. Martin, E. D. Siggia, and H. A. Rose, Phys. Rev. A {\bf 8}, 423 (1973) \bibitem{RaghuPodolskyVishwanath} S. Raghu, D. Podolsky, and A. Vishwanath, to be published. \bibitem{lemberger} K. M. Paget, B. R. Boyce, and T. R. Lemberger, Phys. Rev. B {\bf 59}, 6545 (1999) \bibitem{LeeWen} P.A. Lee and X.G. Wen, Phys. Rev. Lett. {\bf 78}, 4111 (1997) \bibitem{HonerkampLee} C. Honerkamp and P.A. Lee, Phys. Rev. Lett. {\bf 92}, 177002 (2004) \end{thebibliography} \end{document} }
\caption{\label{Ene} Total energy differences per adsorbate atom between different magnetic states of incomplete Cr (left) and Fe (right) layers on W(001) as a function of the coverage: FM state, no relaxation ($+$); DLM state, no relaxation ($\times$); FM state, relaxed (\opensquare) and DLM state, relaxed (\opencircle). The total energy of corresponding NM states is taken as zero.}
\caption{\label{MagMom}Local magnetic moments of Cr (left) and Fe (right) atoms on the W(001) surface as a function of the coverage: unrelaxed FM overlayer ($+$), unrelaxed DLM overlayer ($\times$), relaxed FM overlayer (\opensquare) and relaxed DLM overlayer (\opencircle).}
\caption{MI lobes as a function of $J/U_0$ and $\mu/U_0$, for $|U'|/U_0=0$ (a), $0.25$ (b), $0.5$ (c) and $0.75$. The calculations where made for $L=12$ and $D=5$. The error bars indicate the change in the order parameters, ({\color{red}$\blacksquare$})$\Psi_{PSF}$ and $({\color{blue}{\bullet}})\langle b\rangle$, from $<10^{-4}$ to $>10^{-2}$.}
\caption{({\color{red}$\blacksquare$})$\psi_{PSF}$ and $({\color{green}\bullet})\langle b \rangle$ as a function of $J/U_0$ for $\mu=-0.15$ and $|U'|/U_0=0.75$ with $L=24$ and $D=6$.}
\caption{Amplitude of oscillation in linear response ($V_+=0.03,...,0.07$ V) and hysteretic regime ($V_+=0.09,...,0.17$ V). (\full) stable solutions, (\dashed) unstable solutions. \label{fig:r} }
\caption{Relative phase of oscillations for varying AC-drive amplitude $V_+$. Solid lines indicate stable solutions to (\ref{eq:p1})-(\ref{eq:p3}) while dashed lines correspond to unstable solutions. For small drive amplitudes $V_+\le 0.05$ V the system response agrees with linear response and only stable solutions exist. For intermediate values ($V_+=0.07,...,0.17$ V) hysteresis in the frequency plane appears along high amplitude solutions (mainly unstable). Increasing drive amplitude further a gap in the phase response curve opens up around $\phi=\pi/2$ (curves $V_+=0.19,...,0.27$ V). (\full) stable solutions, (\dashed) unstable solutions. \label{fig:phi} }
\caption{Large amplitude solutions in hysteretic regime ($V_+=0.09,...,0.17$ V). Most of the solutions have oscillation amplitudes too large to be stable. Note, however, that perturbation theory predicts a stable region for the bias $V_+=0.17$ V (red curve). (\full) stable solutions, (\dashed) unstable solutions. \label{fig:r2} }
\caption{Oscillation amplitude $r$ and average displacement $x_0$ in the ``phase gapped'' oscillation regime. Solid lines represent result from perturbation theory for a bias of $V_+=0.23V$ while circles (\opencircle) mark the result of numerical integration of the full system of ODE:s. For small amplitudes the agreement between numerical integration and perturbation theory is very good while perturbation starts to fail for larger amplitude due to the anharmonicity of the tip motion.\label{fig:anal_num}}
\caption{Left: The \jpsi\suppression pattern as measured by NA38 and NA50, in p-A, S-U and Pb-Pb collisions, as a function of L, the distance of nuclear matter traversed by the charmonium, compared to the ``normal nuclear absorption'' band resulting from the Glauber fit to the p-A data~\cite{Goncalo}. The p-A and S-U data points, as well as the curves, have been rescaled to the conditions (energy and rapidity window) of the Pb-Pb data. Right: The suppression of the \psip\as measured by NA38 and NA50 in p-A, S-U and Pb-Pb collisions~\cite{Helena}, compared to the normal nuclear absorption curve calculated using the Glauber formalism~\cite{KLNS}.}
\caption{The \jpsi\and\psip\production cross sections measured in p-A collisions by NA50, as a function of the mass number of the target nuclei (left), and of L (right). The lines represent fits to three commonly used parameterizations of the ``normal nuclear absorption''.}
\caption{NA50 p-A values of the \jpsi\over Drell-Yan cross-section ratio, as a function of A, compared to a Glauber calculation taking into account the feed-down contributions from higher\ccbar\states.}
\caption{The nuclear-dependence $\alpha$ parameter as a function of \xf\(left) and of \pt\(right), for \jpsi\mesons measured at three different collision energies, in p-A collisions~\cite{alpha-pA}.}
\caption{\aptsquared\increases with the size of the target nuclei and with the energy of the collision~\cite{alpha-pA}.}
\caption{\label{fig1}Basic framework for detecting communities as a communication process. A signaler knows the full network structure and wants to send as much information as possible about the network to a receiver over a channel with limited capacity. The signaler therefore encodes the network into modules in a way that maximizes the amount of information about the original network. This figure illustrates an encoder that compresses the network into 3 modules $i=$ {\protect \includegraphics[width=1.7ex]{whitecircle.eps}}, {\protect \includegraphics[width=1.7ex]{blacksquare.eps}}, {\protect \includegraphics[width=1.7ex]{star.eps}}, with $n_i$ nodes and $l_{ii}$ links in each module, and $l_{ij}$ links between the modules. The receiver can then decode the message and construct a set of possible candidates for the original network. The smaller the set of candidates, the more information the signaler has managed to transfer.}
\caption{Information/disturbance trade-off for qudit in a $N$-users transmission line. (Black): $N=1$; (\textcolor{red}{Red}): $N=2$; (\textcolor{green}{Green}): $N=5$; (\textcolor{blue}{Blue}): $N=10$. From left to right the trade-off for $d = \{ 2, 3, 4\}$.}
\caption{Information/disturbance trade-off for qubits in a $2$-user line with different probes' parameter $\theta_A$ and $\theta_B$. In \textcolor{red}{red} the trade-off obtained for $\theta_B=\theta_A$ and varying this single parameter. The black solid line corresponds to the optimal trade-off obtained for $\theta_A=\pi/2$ and varying $\theta_B$. Black dashed curves correspond to trade-offs $F=F_{2,\theta_A}(G)$ obtained varying $\theta_B$ at fixed values of $\theta_A$. From top to bottom $\theta_A=\{4\pi/9 ; \pi/3; 2\pi/9 ; \pi/9 \}$.}