\caption{Energy spectra for elemental groups {\bf a)} protons, {\bf b)} helium, and {\bf c)} iron. Open symbols give results of direct measurements, for references see\Cite{pg,gahbauer,atic03}. Filled symbols represent data from air shower measurements: KASCADE electrons/muons interpreted with two interaction models\Cite{ulrichepj} (preliminary), KASCADE single hadrons\Cite{kascadesh}, and EAS-Top electrons/muons\Cite{eastop-knee}. The data are compared to calculations by Kalmykov \etal\Cite{kalmykov} (\dotted), Sveshnikova\Cite{sveshnikova} (\dashed), and the Poly-Gonato model\Cite{pg} (\line).}
\caption{ Decay chains and \gray\signatures of shortlived radioactive products from explosive SN nucleosynthesis.\label{decay}}{%\footnotesize %\small %\def\arraystretch{0.5} \begin{tabular}{lllrlr} \noalign{\smallskip}\hline\noalign{\smallskip} & & & Decay time & Process &Lines (keV)\\ \noalign{\smallskip} \hline \noalign{\smallskip} $^{56}$Ni $\rightarrow$ & & &8.8 d & EC& 158, 812 \\ &$^{56}$Co $\rightarrow$ &$^{56}$Fe& 111.3 d& EC, $e^+$ (19\%) & 847, 1238\\ %, 2599, 3250\\ \noalign{\smallskip} $^{57}$Ni $\rightarrow$ & & & 52 hr & EC &1370\\ &$^{57}$Co $\rightarrow$ & $^{57}$Fe & 390 d & EC & 122\\ \noalign{\smallskip} $^{44}$Ti $\rightarrow$ & & &86.0 yr & EC &67.9, 78.4\\ &$^{44}$Sc $\rightarrow$ & $^{44}$Ca& 5.7 hr & $e^+$, EC (1\%) &1157\\ \noalign{\smallskip} \hline \end{tabular} }
\caption{The \gray\spectrum of SN 1987A, as accumulated by the\smm-GRS over the period from August 1987 to May 1988 \citep{leising90}. Note that apart from the prominent 847~keV and 1238~keV lines, there is also evidence for the weaker 2599~keV and 3250~keV lines from \cofs. (Courtesy M. Leising)}
\caption{\sax-PDS spectrum of \casa\\citep{vink01a,vink03a}. %The raw count rate spectrum %has been divided by the effective area, in order to obtain aproximate %flux densities. The continuum model consists of a thermal (dash-tripled dotted line) and non-thermal bremsstrahlung model (dashed), the latter based on electron acceleration by lower hybrid wave \citep{laming01b}. The 67.9~keV and 78.4~keV line contributions are indicated jointly by a dotted line. In addition, a potential instrumental background line of Ta-K$\alpha$\is indicated by a dash-dotted line. The observed count-rate spectrum has been divided by the instruments effective area in order to obtain an approximate flux density estimate.\label{pdsspec}}
\caption{ {\red Comparison of the parameter-free LDA+DMFT(QMC) spectra of SrVO$_{3}$ (solid line) and CaVO$_{3}$ (dashed line) with experiments below {\em and} above the Fermi energy. Left panel: high-resolution PES for SrVO$_{3}$ (circles) and CaVO$_{3}$ (rectangles)~\protect\cite{Sekiyama02,Sekiyama03}. Right panel: 1s XAS for SrVO$_{3}$ (diamonds) and Ca$_{0.9}$Sr$_{0.1}$VO$_{3}$ (triangles)~\protect\cite{Inoue94}. Horizontal line: experimental subtraction of the background intensity.} \protect\vspace{-0.3cm} \label{fig_XPS}}
\caption{\label{cap:Scaling_for_queue}Scaling of PR steps $\NPR$ and CPU time $t$ with system size $n$ for FIFO ($\Delta=2.2$). (a) The average number of PR steps per site, $\NPR/n$, as a function of number of push-relabel steps per site per global update, $\Gamma/n$. The data collapse is consistent with the number of operations scaling with $n$ (the best that can be expected for $\xi\ll L$). (b) CPU time $t$ per site ($t/n$) vs.~$\Gamma/n$. The values along both axes have been divided by $n$ the number of sites. \textcolor{magenta}{}A fair collapse develops for $L\ge100$. The minimum is at $\Gamma_{{\rm min}}/n\approx1$, independent of system size. The minimum is shallow, however, and even missing $\Gamma_{{\rm min}}$ by a factor of ten increases the running time only by a factor of about two.}
\caption{\label{exp_calc} Comparison between experimental XANES data (dashed line) and calculated spectra obtained with the \textit{relaxed} models (solid line): a) the Ti $K$ edge in pink sapphire (\pinksapph); b) the Cr $K$ edge in ruby (\ruby); c) the Fe $K$ edge in blue sapphire (\bluesapph). An energy shift was added to valence band maximum in the calculations in order to match with experiments: 4966~eV at the Ti $K$edge, 5991~eV at the Cr $K$ edge and 7118~eV at the Fe $K$ edge.}
\caption{\label{comp_models} Comparison between $\sigma_\parallel$ (left) and $\sigma_\perp$ (right) experimental data (thick solid line) and calculated spectra obtained with both \textit{relaxed} model (thin solid line) and \textit{non-relaxed} model (dashed line): a) the Ti $K$ edge in pink sapphire, \pinksapph; b) the Cr $K$ edge in ruby, \ruby; c) the Fe $K$ edge in blue sapphire, \bluesapph.}
\caption{Experimental superconducting transition temperatures $T_c$ for CeNi$_2$Ge$_2$ ($+$~\cite{Grosche:00}), CePd$_2$Si$_2$ ($\times$~\cite{Grosche:01}), CeIn$_3$ ($*$~\cite{Grosche:01}), CeCoIn$_5$ (\opensquare~\cite{Nicklas:01}, \fullsquare~ \cite{Shishido:03}), CeRhIn$_5$ (\opencircle~\cite{Hegger:00}, \fullcircle~\cite{Muramatsu:01}), plotted as a function of pressure $p$. Lines are guides to the eye. }
\caption{The known/chosen-plaintext attack to multistep parameter modulation, when $10\leq t\leq 30$. Legend: \textcolor{red}{$\Diamond$} -- $0\leq t\leq 10$, $m(t)=1$, $b_s=3.5$; \textcolor{red}{$\ocircle$} -- $10\leq t\leq 20$, $m(t)=1$, $b_s=3.3$; \textcolor{red}{$\Box$} -- $20\leq t\leq 30$, $m(t)=0$, $b_s=3.4$.}
\caption{The known/chosen-plaintext attack to multistep parameter modulation, when $10\leq t\leq 30$. Legend: \textcolor{red}{$\Diamond$} -- $0\leq t\leq 10$, $x_s$-driving, $m(t)=1$, $b_s=3.9$; \textcolor{red}{$\ocircle$} -- $10\leq t\leq 20$, $y_s$-driving, $m(t)=0$, $b_s=3.6$; \textcolor{red}{$\Box$} -- $20\leq t\leq 30$, $x_s$-driving, $m(t)=0$, $b_s=3.2$.}
\caption{ Simulations of star formation in The Mice for (left) density-dependent and (middle) shock-induced star formation recipes; halftones mark old stars, points mark star formation. (Right) Star formation rate vs.\time for density-dependent ({\em solid line}) and various shock-induced ({\em dashed \& dotted}) star formation recipes (from Barnes 2004). }
\caption{(color online). QMC (points) and DMRG (continuous lines) results for $E_b(10)=E(5,5)+E(6,6)-2E(5,6)$ vs $U/t$ for: $V=7.5t$ (\textcolor{red}{$\bigtriangledown$},\textcolor{green}{$\bigcirc$}), and $V=9.5t$ (\textcolor{cyan}{$\Diamond$},\textcolor{blue}{$\triangle$}) (see I), when the center of the trap is located in the middle of two lattice points (\textcolor{red}{$\bigtriangledown$},\textcolor{cyan}{$\Diamond$}), and in a lattice point (\textcolor{green}{$\bigcirc$},\textcolor{blue}{$\triangle$}). The dashed line shows ED results for $E_b(10)$ in open MI systems (see text). The inset shows ED results for $E^*_G=E_G+AN_f$ vs $N_f$, in open MI systems with $U/t=8$. Here, we choose a nonzero value of $A=0.327$ to stress the even(\textcolor{green}{$\times$})-odd(\textcolor{red}{$+$}) effect, without affecting the actual value of $E_b<0$ it causes.}
\caption{MODEL 2: Face-on view of a N-body model of a barred galaxy with a weak bar and a highly concentrated center (pseudobulge) The plot shows contours of equal surface density. The circle represents the corotation radius for this model. }{\includegraphics[width =0.47\textwidth]{Avcontour.8770.ps}}
\caption{Distribution of particles in the frequency space for model 2. Horizontal axis shows angular frequencies $(\Omega)$ and vertical axis shows radial frequencies $(\kappa)$. Each point represents an orbit of a particle during 1 Gyr of evolution. Clustering of particles along straight lines with certain slopes indicates the presence of resonances. }{\includegraphics[width =0.47\textwidth]{mapping.1G.3.2.ps}}
\caption{Spatial distribution of particles at the inner Linblad resonance in the model 2. The distribution is not localized at any radius, but it is spread along the bar. ILR particles can come arbitrary close to the center. The circle shows the corotation radius. The bar stays parallel to the horizontal axis. }{\includegraphics[width =0.47\textwidth]{freposi.OL.ps}}
\caption{ Spatial distribution of particles at the ultra-harmonic resonance (-1:4) in the model 2. The circle presents the corotation radius. }{\includegraphics[width =0.47\textwidth]{freposi.9.ps}}
\caption{Spatial distribution of particles at the 1:5 resonance in the model 2. The circle presents the corotation radius. }{\includegraphics[width =0.47\textwidth]{DprofAv.G3.4.ps}}
\caption{ Evolution of the spatial distribution of particles, which were at the corotation resonance at 3.5 Gyr for model 3. The circle represent the corotation radius. Particles trapped near corotation stay there for a long time. }{\includegraphics[width =0.47\textwidth]{AMmapping.8750.2.ps}}
\caption{Example of a trajectory near the corotation resonance. The particle circulates around the corotation radius (circle) after it got trapped. The reference frame rotates with the pattern speed. So, the bar stays always along the horizontal axis. }{\includegraphics[width =0.47\textwidth]{trayelist9C1.1.ps}}
\caption{ Example of libration around the corotation radius. The particle librates around the areas of maximum positive and negative change of the angular momentum.}{\includegraphics[width =0.47\textwidth]{trayelist4G2.1.ps}}
\caption{\label{Fig:UndoParticipants}% Participants in \jhotdraw's undo implementation.}
\caption{ \small The distribution of remaining events over the $\pm 10 \sigma $ region in the $ \Minv - \Delta E $ plane after opening the blinded region. The ellipse is the signal region that has a signal acceptance of $\Omega = 90\%$. The blinded regions $\pm 5\sigma$ in $\Minv$ and $\pm 0.5$ GeV in $\Delta E$ are indicated by the dotted lines. The cross indicates the $\pm 2\sigma$ $\Minv$ and $\Delta E$ ranges. Various symbols show events from 428 fb$^{-1}$ of generic $\tau^+\tau^-$ MC (\textcolor{blue}{$\Box$}), 232 fb$^{-1}$ of $q\bar{q}$ MC with $q = u,d,s$ (\textcolor{red}{$\vartriangle$}), 160 fb$^{-1}$ with $q = c$ (\textcolor{red}{$\circ$}), 205 fb$^{-1}$ of two-photon MC (\textcolor{black}{$\blacktriangledown$}), and data (\textcolor{black}{$\bullet$}). }
\caption{ Matching field-dependence of the temperature derivative of the characteristic fields, $dH_{Cp}(T)/dT$ and $dH_{irr}(T)/dT$, in single crystalline YBa$_{2}$Cu$_{3}$O$_{7-\delta}$ ( \color{yellow} $\triangle$ \color{black}, \color{magenta}$\Box$\color{black}, \color{red} \Large $\circ$ \normalsize \color{black}) and K$_{0.35}$Ba$_{0.65}$BiO$_{3}$ ($\triangle$); Whereas $dH_{irr}(T)/dT$ and $dH_{Cp}(T)/dT$ coincide for K$_{x}$Ba$_{1-x}$BiO$_{3}$, the two lines are well separated for YBa$_{2}$Cu$_{3}$O$_{7-\delta}$. Also, in contrast to the results of Ref.~\protect\onlinecite{Samoilov96} ( \color{magenta}$\Box$\color{black}), no saturation of $dH_{irr}(T)/dT$ is found in YBa$_{2}$Cu$_{3}$O$_{7-\delta}$. %Closed symbols (\protect\rule{1.5mm}{1.5mm}) and ($\bullet$) denote %the locus of $|F_{s}(B)+n_{t}U_{p}(B)-F_{n}| = %\alpha k_{B}T/V$ (see text). }
\caption{\label{fig:example}A few samples from the Monte Carlo used to produce \tref{tab:example1} are displayed in detail. They refer to $N_s=5$, and the background is (\textit{a}) $N_b=50$ (\textit{b}) $N_b=0.5$ (\textit{c}) $N_b=0.01$. In the plots above the cumulative histogram of the $p$-values is compared with the threshold given by the Bonferroni (\broken) and the BH (\dashed) procedures.}
\caption{The plot shows the model potential with the ground-state (\full), the first excited state (\dashed), the second excited state (\dotted) and the third excited state (\chain). Each state is shifted according to its eigenvalue.}
\caption{Simulation results showing formation of clusters of sphingolipids (\textcolor[rgb]{0,0,1}{*}) \& cholesterol(\textcolor[rgb]{0,1,0}{x}) at different cholesterol concentrations. The membrane configurations demonstrate formation of domains in the cholesterol range $0 < x_c \le 0.2$.}
\caption{ The 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 criterium (\ref{criterium}) 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{ The results of $\Delta u_V$ and $\Delta d_V$ reconstruction in the region $[a_{min}=10^{-4},b_{max}=1]$ 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 corresponds to the curve reconstructed in the entire $[a_{\rm max}=10^{-4},b_{\rm max}=1]$ region with requirement of minimal deviation from the curve (bold solid line ) reconstructed with MJEM and criterium (\ref{criterium}) inside the accessible for measurement region ([0.023,0.6] here). }{\includegraphics[height=4cm,width=6cm]{new-tail-uv-sym.eps}}{\includegraphics[height=4cm,width=6cm]{new-tail-dv-sym.eps}}{\includegraphics[height=4cm,width=6cm]{new-tail-uv-br.eps}}{\includegraphics[height=4cm,width=6cm]{tail-dv-br-final.eps}}
\caption{Radiative correction to $\delta$, defined as $\Delta_\mathrm{RG}\delta\equiv \big|\delta(m_Z)- \delta(M_\mathrm{GUT})\big|$ for (a) $\theta_{13}(m_Z)=10^\circ$ and (b) $\theta_{13}(m_Z)=0.1^\circ$ as a function of $\tan\beta$ and $m_1$. We use $\delta(m_Z)=90^\circ$.} \label{fig:DeltaRGdelta} \end{figure} Let us finally mention that RG effects for Dirac neutrinos will always result in a rescaling of the mass eigenvalues. Beyond that, in the framework of the SM, mixing parameters are quite stable. The only exceptions are $\theta_{12}$ for very degenerate masses, and $\delta$ for tiny $\theta_{13}$. On the other hand, in the MSSM, RG effects are increased by $\tan^2\beta$, i.e.\ by up to three orders of magnitude. \section{Summary} Assuming Dirac neutrinos, we have derived renormalization group equations for leptonic mixing parameters. The results share several features with the corresponding equations for Majorana neutrinos. However, Dirac running is more predictive, as the Majorana phases are unphysical in this case. This makes it possible to specify the amount of renormalization group evolution unambiguously as soon as $m_1$ and $\delta$ (and $\tan\beta$) are known. The renormalization group evolution alone does not yield an explanation of the largeness of the leptonic mixing angles (for an analogous and very clear discussion for Majorana neutrinos see \cite{Casas:2003kh}). Yet it may account for radiative enhancement of $\theta_{12}$, and possibly also of $\theta_{23}$, since both can increase significantly in the MSSM when running down. Most importantly, we find that in phenomenological studies renormalization group effects for leptonic mixing angles can in general not be neglected. This can be understood from the fact that $\Dot{\theta}_{ij} = {f(m_i,\theta_{ij},\delta)}/({m_i^2-m_j^2})$ which becomes singular if $m_i\to m_j$ and vanishes if the mixing angles are zero. We have thus traced back the relative enhancement of the quantum corrections of leptonic mixing parameters as compared to quark mixings to two reasons. First, the mass hierarchy which suppresses the renormalization group running, can be much weaker. Second, the mixing angles are larger so that the parameters are further apart from the renormalization group stable situation where all mixings are zero. As there is no suppression of the running by phases, the renormalization group corrections should in general be taken into account even for a strong hierarchy to accommodate the precision of future experiments. Renormalization group corrections are especially relevant if the mass spectrum is non-hierarchical, and $\tan\beta$ is large in the MSSM. Hence, similarly to the case of Majorana neutrinos \cite{Antusch:2003kp}, also in the Dirac case the non-observation of deviations of the angles from special points, e.g.\ of $\theta_{12}$ from $\pi/4-\vartheta_{12}$ (with $\vartheta_{12}$ being the Cabibbo angle), of $\theta_{13}$ from 0 and $\theta_{23}$ from $\pi/4$, may restrict the parameters such as the absolute neutrino mass scale. The current experimental data already has the necessary precision to indicate disfavored parameter regions. It may also point to exactly realized symmetries and our formulae can hence be used to identify possible symmetries. Whenever a symmetry is exact and fixes some mixing parameters, those mixing parameters have to be stable under quantum corrections. For instance, it has recently been pointed out \cite{Grimus:2004cj} that for Majorana neutrinos and an inverted hierarchy the configuration $m_3=\theta_{13}=0$ is stable. From the analytic expressions it is obvious that this statement also applies to the Dirac case. Likewise, a quick inspection of the RGEs excludes most of the proposed symmetries from being exact. Our formulae are basis-independent, and thus allow to understand certain features of the underlying theory, such as symmetries, in a basis-independent way. We have discussed this only for the case of CP symmetry, but it is obvious how the analysis can be carried over to other symmetries. In this context, it would be interesting to see if infrared fixed points with large mixings, as discussed in \cite{Casas:2002sn,Casas:2003kh}, can be obtained for (non-standard) Dirac neutrinos as well \cite{Dedes:prep}. In this case, one may hope for a scenario where the large mixings are a consequence of running, and the mechanism of generation of neutrino masses is still related to the scale where gauge couplings meet. We conclude that in the light of future precision experiments, flavor physics might enter into an era of ``precision model building''. It seems possible to determine the mixing parameters to a remarkable accuracy, precise enough such that flavor models and the corresponding renormalization group effects become to a certain degree distinguishable. For a specific parameter and a desired accuracy, our formulae allow to estimate the renormalization group effects, and to judge to which extent a numerical analysis is required. \subsubsection*{Acknowledgements} We would like to thank S.~Antusch, A.~Dedes and J.~Kersten for valuable discussions. One of us (M.R.) would like to thank the Aspen Center for Physics for support, and the CERN theory group for hospitality. This work was partially supported by the EU 6th Framework Program MRTN-CT-2004-503369 ``Quest for Unification'' and MRTN-CT-2004-005104 ``ForcesUniverse''. This work was also supported by the Deutsche Forschungsgemeinschaft in the ``Sonderforschungsbereich 375 f\"ur Astro-Teilchenphysik'' and under project number RO-2516/3-1. \vspace*{1cm} \appendix \renewcommand{\thesection}{\Alph{section}} \renewcommand{\thesubsection}{\Alph{section}.\arabic{subsection}} \def\theequation{\Alph{section}.\arabic{equation}} \renewcommand{\thetable}{\arabic{table}} \renewcommand{\thefigure}{\arabic{figure}} \setcounter{section}{0} \setcounter{equation}{0} \section{Mixing parameters RGEs for Dirac masses} \subsection{Lagrangian} The Yukawa couplings are given by \begin{equation} -\mathcal{L}_\mathrm{Yuk}=\left(Y_\nu\right)_{gf}\overline{N_R^g}\tilde{\phi}^\dagger\ell_L^f+\left(Y_e\right)_{gf}\overline{e_R^g}\phi^\dagger\ell_L^f+\left(Y_u\right)_{gf}\overline{u_R^g} \tilde{\phi}^\dagger Q_L^f+\left(Y_d\right)_{gf}\overline{d_R^g} \phi^\dagger Q_L^f \end{equation} in the SM extended by right-handed neutrinos where $\tilde{\phi}=i\sigma_2\phi^*$. In the extended MSSM, the Yukawa couplings are analogously defined in the superpotential by \begin{equation} \mathcal{W}_\mathrm{Yuk}= \left(Y_\nu\right)_{gf} N_R^{Cg}\phi^{(2)}\epsilon^T\ell_L^f + \left(Y_e\right)_{gf} e_R^{Cg} \phi^{(1)}\epsilon\ell_L^f + \left(Y_u\right)_{gf} u_R^{Cg}\phi^{(2)}\epsilon^T Q_L^f+\left(Y_d\right)_{gf}d_R^{Cg}\phi^{(1)}\epsilon Q_L^f\; . \end{equation} The left-handed lepton and quark doublets are denoted by $\ell_L$ and $Q_L$, respectively. We assume that there is no Majorana mass term for the right-handed neutrinos. \subsection{$\boldsymbol{\beta}$-functions} The relevant $\beta$-functions for the down-type quark, up-type quark, charged lepton and neutrino Yukawa coupling matrices $Y_d$, $Y_u$, $Y_e$ and $Y_\nu$ read at one-loop~\cite{Machacek:1984fi,Grzadkowski:1987tf} \begin{subequations} \begin{eqnarray} (4\pi)^2\Dot{Y}_d & = & Y_d\left\{ C_d^d\,Y_d^\dagger Y_d +C_d^u\, Y_u^\dagger Y_u +\alpha_d \right\}\;, \\ (4\pi)^2 \Dot{Y}_u & = & Y_u\left\{ C_u^d\,Y_d^\dagger Y_d + C_u^u\, Y_u^\dagger Y_u + \alpha_u \right\}\;, \\ (4\pi)^2 \Dot{Y}_e & = & Y_e\left\{ C_e^e\,Y_e^\dagger Y_e +C_e^\nu\, Y_\nu^\dagger Y_\nu +\alpha_\ell \right\}\;, \\ (4\pi)^2 \Dot{Y}_\nu & = & Y_\nu\left\{ C_\nu^e\,Y_e^\dagger Y_e + C_\nu^\nu\, Y_\nu^\dagger Y_\nu + \alpha_\nu \right\}\;,\label{eq:DiracRGE} \end{eqnarray} \end{subequations} where \begin{subequations} \begin{align} C_d^d & = \left\{\begin{array}{ll} 3/2\;, & \text{(SM)}\\ 3\;, & \text{(MSSM)} \end{array}\right. & C_d^u & = \left\{\begin{array}{ll} -3/2\;, & \text{(SM)}\\ 1\;, & \text{(MSSM)} \end{array}\right.\\ C_u^d & = \left\{\begin{array}{ll} -3/2\;, & \text{(SM)}\\ 1\;, & \text{(MSSM)} \end{array}\right. & C_u^u & = \left\{\begin{array}{ll} 3/2\;, & \text{(SM)}\\ 3\;, & \text{(MSSM)} \end{array}\right. \\ C_e^e & = \left\{\begin{array}{ll} 3/2\;, & \text{(SM)}\\ 3\;, & \text{(MSSM)} \end{array}\right. & C_e^\nu & = \left\{\begin{array}{ll} -3/2\;, & \text{(SM)}\\ 1\;, & \text{(MSSM)} \end{array}\right.\\ C_\nu^e & = \left\{\begin{array}{ll} -3/2\;, & \text{(SM)}\\ 1\;, & \text{(MSSM)} \end{array}\right. & C_\nu^\nu & = \left\{\begin{array}{ll} 3/2\;, & \text{(SM)}\\ 3\;, & \text{(MSSM)} \end{array}\right. \end{align} \end{subequations} and \begin{subequations} \begin{eqnarray} \alpha_d & = & \left\{\begin{array}{ll} - \frac{1}{4} g_1^2 - \frac{9}{4} g_2^2 - 8\,g_3^2 + T_\mathrm{SM}\;, & \text{(SM)}\\[0.2cm] 3\,\Tr(Y_d^\dagger Y_d) + \Tr(Y_e^\dagger Y_e) - \frac{7}{15}\,g_1^2 - 3\,g_2^2 - \frac{16}{3}\,g_3^2\;, & \text{(MSSM)} \end{array}\right. \\[0.2cm] \alpha_u & = & \left\{\begin{array}{ll} - \frac{17}{20} g_1^2 - \frac{9}{4} g_2^2 - 8\,g_3^2 + T_\mathrm{SM}\;, & \text{(SM)}\\[0.2cm] \Tr( Y_\nu ^\dagger Y_\nu ) + 3\,\Tr(Y_u^\dagger Y_u) - \frac{13}{15}\,g_1^2 - 3\,g_2^2 - \frac{16}{3}\,g_3^2 \hphantom{- \frac{16}{3}\,g_3^2}\;, & \text{(MSSM)} \end{array}\right. \\ \alpha_\ell & = & \left\{\begin{array}{ll} - \frac{9}{4} g_1^2 - \frac{9}{4} g_2^2 + T_\mathrm{SM}\;, & \text{(SM)}\\[0.2cm] 3\,\Tr(Y_e^\dagger Y_e) + \Tr(Y_\nu^\dagger Y_\nu) - \frac{9}{5}\,g_1^2 - 3\,g_2^2 \hphantom{- \frac{16}{3}\,g_3^2}\;, & \text{(MSSM)} \end{array}\right. \\[0.2cm] \alpha_\nu & = & \left\{\begin{array}{ll} - \frac{9}{20}\,g_1^2 - \frac{9}{4}\,g_2^2 + T_\mathrm{SM}\;, & \text{(SM)}\\[0.2cm] \Tr( Y_\nu ^\dagger Y_\nu ) + 3\,\Tr(Y_u^\dagger Y_u) - \frac{3}{5}\,g_1^2 - 3\,g_2^2 \;, & \text{(MSSM)}\;. \end{array}\right. \label{eq:AlphaNu} \end{eqnarray} \end{subequations} Here, we define $T_\mathrm{SM}\equiv\Tr\left[Y_e^\dagger Y_e + Y_\nu^\dagger Y_\nu + 3\,Y_d^\dagger Y_d + 3\,Y_u^\dagger Y_u\right]$, and use GUT normalization for $g_1$. \subsection{General derivation} \label{subsec:MixingParameterRGEsDirac} In this subsection, we will perform a general analysis applicable for any Dirac masses, and treat the evolution of lepton and quark masses and mixings only as a special case. We derive the running of mixing parameters for a RGE of the form \begin{equation} 16\pi^2\,\frac{\D}{\D t} H ~=~ F^\dagger\,H+H\,F + f\, H\;,\end{equation} where $f$ is real and $H$ is Hermitean, so that we can diagonalize it in a `reference basis', \begin{equation} U^\dagger\cdot H\cdot U ~=~ D\;.\end{equation} In the application in the main part, $F$ corresponds either to $C_d^u\,Y_u^\dagger Y_u+C_d^dY_d^\dagger Y_d$ (or $C_\nu^e\,Y_e^\dagger Y_e+C_\nu^\nu Y_\nu^\dagger Y_\nu$ for the lepton sector), and $H$ to $Y_d^\dagger Y_d$ (or $Y_\nu^\dagger Y_\nu$). The `reference basis' is the basis where $Y_u^\dagger Y_u$ (or $Y_e^\dagger Y_e$) is diagonal at $t=t_0$. $U$ denotes then to the CKM matrix $U_\mathrm{CKM}$ (or the MNS matrix $U_\mathrm{MNS}$). $f$ denotes the diagonal parts of the $\beta$-function, $f=2\alpha_d$ (or $f=2\alpha_\nu$). Now we perform an analysis very similar to what is done in \cite{Antusch:2003kp} which is based on \cite{Grzadkowski:1987tf,Babu:1987im,Casas:1999tg}. We can differentiate the relation $H=U\cdot D\cdot U^\dagger$, \begin{eqnarray} \frac{\D}{\D t}(U\cdot D\cdot U^\dagger) & = & \Dot U\cdot D\cdot U^\dagger + U\cdot D\cdot \Dot U^\dagger + U\cdot \Dot D\cdot U^\dagger \nonumber\\ & \stackrel{!}{=} & \frac{1}{16\pi^2} \left(F^\dagger \cdot U \cdot D \cdot U^\dagger +U\cdot D\cdot U^\dagger \cdot F + f\,U\cdot D \cdot U^\dagger\right) \;.\end{eqnarray} Multiplying by $U^\dagger$ from the left and by $U$ from the right yields \begin{equation} U^\dagger\cdot \Dot U \cdot D + D\cdot \Dot U^\dagger\cdot U + \Dot D ~=~\frac{1}{16\pi^2} \left(F^{\prime\,\dagger}\cdot D + D\cdot F' +f\,D\right) \;,\end{equation} where $F'=U^\dagger\cdot F\cdot U$. For the quark case, $F'=C_d^d\,D+C_d^u\,U^\dagger Y_u^\dagger Y_u\,U$. We will see below that only the off-diagonal components are relevant for the RGEs of the mixing parameters. The evolution of $U$ can be written as \begin{equation}\label{eq:EvolutionOfU} \frac{\D}{\D t} U~=~U\cdot X\;,\end{equation} where $X$ is anti-Hermitean. Inserting this relation yields \begin{equation} \Dot D +X\cdot D + D\cdot X^\dagger ~=~ \frac{1}{16\pi^2}\left(F^{\prime\,\dagger}\cdot D + D\cdot F' +f\,D\right) \;,\end{equation} or, by using the anti-Hermitecity of $X$, \begin{equation} \Dot D ~=~ \frac{1}{16\pi^2}\left(f\,D +F^{\prime\,\dagger}\cdot D + D\cdot F'\right) -X\cdot D + D\cdot X\;.\end{equation} Denoting the entries of $D$ by $y_i^2$, i.e.\$D=\diag(y_1^2,y_2^2,y_3^2)$, we find \begin{equation} \frac{\D}{\D t}y_i^2 ~=~ \frac{1}{16\pi^2}\left[f\,y_i^2+(F_{ii}^{\prime\,*}+F_{ii}')\,y_i^2\right] \;,\end{equation} i.e.\the terms proportional to$X$ have dropped out. This equation corresponds to a RGE for the running mass eigenvalues, defined by $m_i(t)=|y_i(t)|\,v$ with $v$ fixed, of the form \begin{equation} (4\pi)^2\Dot{m}_i ~=~ (\re F_{ii}'+\alpha)\,m_i\;.\end{equation} By analyzing the off-diagonal parts we obtain \begin{equation} y_i^2\,X_{ij} - X_{ij}\,y_j^2 ~=~ -\frac{1}{16\pi^2}\left[(F^{\prime\,\dagger})_{ij}\,y_j^2 +y_i^2\,F'_{ij}\right] \;.\end{equation} This can be converted into equations for real and imaginary part of $X$, which, since $F$ is Hermitean, can be combined to \begin{equation} \label{eq:XfromFprime} 16\pi^2 X_{ij} = \frac{y_j^2+y_i^2}{y_j^2-y_i^2} \, F_{ij}' \;.\end{equation} The diagonal parts of $X$ remain undetermined. However, this is not a problem, since they only influence the RG evolution of the unphysical phases.\footnote{Note that the Majorana phases are unphysical in the the Dirac case as well.} So far, we have derived the differential change of the CKM matrix due to the RG corrections for $Y_d^\dagger Y_d$ (cf.\Eq.~\eqref{eq:EvolutionOfU}). In the Majorana neutrino case, the analogous differential equation already describes the evolution of the MNS matrix since $Y_e^\dagger Y_e$ doesn't get rotated by the RGE.\footnote{This is only true at leading order in $M^{-1}$ where $M$ denotes the scale of the effective neutrino mass operator (e.g.\the see-saw scale)\cite{Broncano:2004tz}.} For Dirac neutrinos, $Y_e^\dagger Y_e$ gets rotated only by terms proportional to the squares of Dirac Yukawa couplings, hence those rotations can safely be neglected. In the quark sector, the radiative rotation of $Y_u^\dagger Y_u$ represents an important effect, as we will argue in the following. \subsection{Contribution from the change of $\boldsymbol{Y_u}$} Here, we specialize to the quark sector as the analogous effect is irrelevant for Dirac neutrinos. The RGE for $Y_u$ contains non-diagonal terms so that continuous re-diagonalization is required. Since the mixing matrix $U_\mathrm{CKM}$ is defined as the matrix which diagonalizes $Y_d^\dagger Y_d$ in the basis in which $Y_u$ is diagonal, $U_\mathrm{CKM}$ receives an additional contribution from the running of $Y_u$, \begin{equation} \frac{\D}{\D t}U_\mathrm{CKM} ~=~ U_\mathrm{CKM}\cdot X+ \text{term stemming from the change of}\:Y_u\;.\end{equation} To evaluate this change, we can essentially repeat the steps of the previous subsection. Introducing a matrix $\widetilde{U}$ which diagonalizes $Y_u^\dagger Y_u$ in the reference basis (implying $\widetilde{U}(t=t_0)=\mathbbm{1}$), i.e.\\begin{equation} \widetilde{U}^\dagger\,Y_u^\dagger Y_u\,\widetilde{U} ~=~ \diag(\widetilde{y}_1^2,\widetilde{y}_2^2,\widetilde{y}_3^2)\;,\end{equation} we arrive at \begin{equation} \frac{\D}{\D t}\widetilde{U} ~=~ \widetilde{U}\cdot \widetilde{X}\;,\end{equation} where the off-diagonal entries of $\widetilde{X}$ are given by \begin{equation}\label{eq:Xtilde} 16\pi^2\,\widetilde{X}_{ij} ~=~ \frac{\widetilde{y}_i^2+\widetilde{y}_j^2}{\widetilde{y}_j^2-\widetilde{y}_i^2} \,\widetilde{F}_{ij}\;.\end{equation} Completely analogous to \ref{subsec:MixingParameterRGEsDirac}, \begin{equation} \widetilde{F}' ~=~ \widetilde{U}^\dagger\cdot\widetilde{F}\cdot\widetilde{U}\;,\end{equation} and at $t=t_0$ \begin{equation} \widetilde{F}' ~=~C_u^d\,U\,D\,U^\dagger+C_u^u\,Y_u^\dagger Y_u\;.\end{equation} Again, only the off-diagonal terms influence the RGEs of the mixing angles. \subsection{Mixing parameter RGEs in the quark sector} As $U_\mathrm{CKM}=\widetilde{U}^{-1}U=\widetilde{U}^\dagger U$, the RGE for the CKM matrix reads \begin{equation} \frac{\D}{\D t}U_\mathrm{CKM} ~=~ \widetilde{X}^\dagger\cdot U_\mathrm{CKM}+U_\mathrm{CKM}\cdot X\;.\end{equation} To proceed, we label the mixing parameters by \begin{equation} \{\xi_k\} = \{\theta_{12},\theta_{13},\theta_{23}, \delta,\delta_e,\delta_\mu,\delta_\tau, \varphi_1,\varphi_2\} \;,\end{equation} and evaluate the derivative of $U_\mathrm{CKM}$, \begin{equation} \Dot U_\mathrm{CKM} ~=~ \Dot U_\mathrm{CKM} \left(\{\Dot \xi_k\},\left\{\xi_k\right\}\right) \;.\end{equation} Observe that the resulting expression is linear in $\Dot\xi_k$. By solving a system of linear equations of the form \begin{equation} \sum\limits_k A^{(k)}_{TX}\,\Dot\xi_k + \I\,S^{(k)}_{TX}\,\Dot\xi_k ~=~ R_{X}\;,\end{equation} where \begin{equation} R_{X} ~=~ U_\mathrm{MNS}\cdot T + X^\dagger \cdot U_\mathrm{MNS}\;,\end{equation} we thus obtain a set of linear equations for the $\Dot\xi_k$. RGEs for the matrix elements have been derived in refs.~\cite{Babu:1987im,Grzadkowski:1987tf}. From these, we obtain the RGEs for the mixing angles in the quark sector. Neglecting all Yukawa coupling except for $y_t$ and $y_b$, they read \begin{subequations} \begin{eqnarray} \Dot{\vartheta}_{12} & = & \frac{C_{d}^{u}\,{{y_t}}^2}{64\,{\pi }^2} \,\cos ({{\vartheta }_{12}})\,\left\{\left[ \left(3 - \cos 2\,\vartheta_{13} \right)\,\cos 2\,\vartheta_{23} - 2\,\cos^2\vartheta _{13} \right]\,\sin \vartheta_{12} \right.\nonumber\\ & & \qquad\left.{} + 4\,\cos \delta_\mathrm{CP} \,\cos \vartheta_{12}\,\sin \vartheta_{13}\,\sin 2\,\vartheta _{23} \right\} \;,\\ \Dot{\vartheta}_{13} & = & \frac{-\sin 2\,\vartheta_{13}}{64\,{\pi }^2} \left[2\,C_{u}^{d}\,y_b^2 + C_{d}^{u}\,y_t^2\,\left(1+ \cos 2\,\vartheta_{23}\right)\right] \;,\\ \Dot{\vartheta}_{23} & = & \frac{-\sin 2\,\vartheta_{23}}{64\,{\pi }^2} \left[C_{u}^{d}\,y_b^2\,\left(1+\cos 2\,\vartheta_{13}\right) + 2\,C_{d}^{u}\,y_t^2\right] \;.\end{eqnarray} \end{subequations} It turns out that finite $y_s$ and $y_c$ corrections yield an important but sub-dominant effect for $\Dot{\vartheta}_{12}$. The dominant term in the RGE of $\delta_\mathrm{CP}$ is \begin{equation} \Dot{\delta}_\mathrm{CP}~=~ \frac{C_{d}^{u}\,y_s^2\,y_t^2}{8\,\pi^2\,\left(y_b^2 - y_s^2 \right)} \,\cos \vartheta_{12}\,\cos \vartheta_{23}\,\sin\delta\,\sin \vartheta_{12}\,\sin \vartheta_{23}\times\vartheta_{13}^{-1}\;.\end{equation} \subsection{Mixing parameter RGEs in the (Dirac) neutrino sector} In order to derive analogous RGEs for the leptonic mixing parameters, observe that the RG change of $Y_e^\dagger Y_e$ is quadratic in neutrino Yukawa couplings, i.e.\strongly suppressed. Thus, we can safely neglect the$\widetilde{X}$ contribution, \begin{equation} \frac{\D}{\D t}U_\mathrm{MNS} ~=~ \widetilde{X}^\dagger\cdot U_\mathrm{MNS}+U_\mathrm{MNS}\cdot X ~\simeq~ U_\mathrm{MNS}\cdot X\;,\end{equation} where $X$ is now related to $F'$ by Eq.~\eqref{eq:XfromFprime}, and $F'=C_\nu^\nu\,D+C_\nu^e\,U_\mathrm{MNS}^\dagger Y_e^\dagger Y_e\,U_\mathrm{MNS}$ at $t=t_0$. \bibliography{Running} \bibliographystyle{ArXiv} \end{document} }\end{equation}}
\caption{Yeast protein interaction data (o) and best-fit probability distributions: Poisson (\textcolor{black}{\bf ---}), Exponential (\textcolor{yellow}{\bf ---}), Gamma (\textcolor{cyan}{\bf ---}), Power-law (\textcolor{red}{\bf ---}), Lognormal (\textcolor{green}{\bf ---}), Stretched exponential (\textcolor{blue}{\bf ---}). The parameters of the distributions shown in this figure are the maximum likelihood estimates based on the real observed data. }
\caption{Degree distributions of the protein interaction networks (o) of {\it C.elegans}, {\it D.melanogaster}, {\it E.coli} and {\it H.pylori}. The power-law (\textcolor{red}{\bf ---}), log-normal (\textcolor{green}{\bf ---}) and stretched exponential (\textcolor{blue}{\bf ---}) models are shown for all figures; for {\it E.coli} the gamma distribution (\textcolor{cyan}{\bf ---}), which performs better (measured by the Akaike weights) than either scale-free and the stretched exponential distributions.}
\caption{Degree distributions of the metabolic network data (o) and best-fit probability distributions: Poisson (\textcolor{black}{\bf ---}), Exponential (\textcolor{yellow}{\bf ---}), Gamma (\textcolor{cyan}{\bf ---}), Power-law (\textcolor{red}{\bf ---}), Log-normal (\textcolor{green}{\bf ---}), Stretched exponential (\textcolor{blue}{\bf ---}). The parameters of the distributions shown in this figure are the maximum likelihood estimates based on the real observed data. Ignorig low degrees ($k=1,2$) when fitting the scale-free model increases the exponent (\ie it falls off more steeply) but compared to the other models no increased performance is observed.}
\caption{A presentation of the globally-integrated 1-850\m\spectral energy distributions for 10 SINGS galaxies.\2MASS,\Spitzer, \IRAS, \ISO, and \SCUBA\data are represented by filled squares, filled circles, filled triangles, open circles, and open squares, respectively. The solid curve is the sum of a dust (dashed) and a stellar (dotted) model. The dust curve is a Dale\& Helou (2002) model fitted to the 24, 70, and 160\m\fluxes; the$\alpha_{\rm SED}$ listed within each panel parametrizes the distribution of dust mass as a function of heating intensity, as described in Equation~\ref{eq:dMdU} and Dale \& Helou (2002). The stellar curve is the 900~Myr continuous star formation, solar metallicity, Salpeter IMF ($\alpha_{\rm IMF}=2.35$) curve from Vazquez \& Leitherer (2005) fitted to the 2MASS data.}
\caption{A representative collection of empirical SEDs from the SINGS sample. Normalizing at 8.0\m\shows the dramatic variations over infrared wavelengths, especially in the near-infrared.}
\caption{Top: The {\it Spitzer} (dust-only) mid- and far-infrared color-color diagram for globally-integrated SINGS data. Filled symbols represent low metallicity galaxies. The solid and dotted lines indicate the dust-only SED models of Dale \& Helou (2002) and Dale et al. (2001), respectively, derived from the average global trends for a sample of normal star-forming galaxies observed by\ISO\and\IRAS. Bottom: The 8.0\m\(dust-only) flux with respect to the total 3-1100\m\dust emission, as a function of far-infrared color. Note that only low-metallicity galaxies are significantly below the normal star-forming galaxy dust curves (3$\sigma$ upper limits are provided for DDO~154). A typical set of error bars are provided in both panels for reference.}
\caption{The ratio of dust-only 8.0\m\emission to the total 3-1100\m\dust emission as a function of the far-infrared color. The relative constancy in the ratio is consistent with the interpretation that elevated 24\m\emission (not reduced PAH emission) drives the\Spitzercolorc\ratio to low values in regions of higher dust temperature in these three galaxies.}
\caption{Top: The range of {\it Spitzer} \Spitzercolorf\colors as a function of redshift, predicted for systems similar to SINGS galaxies. Bottom: The (dust) bolometric infrared flux, scaled to the observed 24\m\flux, as a function of redshift.}
\caption{Size of the central peak after 20~ms of ballistic expansion versus the number of atoms. The clouds were released from the following potentials MT (red \textcolor{red}{$\circ$}), MT plus DP (black $\Box$), MT plus OL (blue \textcolor{blue}{$\lhd$}), MT plus DP plus OL (green \textcolor{green}{$\rhd$}). The lines correspond to a theoretical prediction (see text). The lattice depth was 6.5~$E_r$ and the DP had a depth of 0.1~$E_r$.}
\caption{Yeast protein interaction data (o) and best-fit probability distributions: Poisson (\textcolor{black}{\bf ---}), Exponential (\textcolor{yellow}{\bf ---}), Gamma (\textcolor{cyan}{\bf ---}), Power-law (\textcolor{red}{\bf ---}), Lognormal (\textcolor{green}{\bf ---}), Stretched exponential (\textcolor{blue}{\bf ---}). The parameters of the distributions shown in this figure are the maximum likelihood estimates based on the real observed data. }
\caption{Degree distributions of the protein interaction networks (o) of {\it C.elegans}, {\it D.melanogaster}, {\it E.coli} and {\it H.pylori}. The power-law (\textcolor{red}{\bf ---}), lognormal (\textcolor{green}{\bf ---}) and stretched exponential (\textcolor{blue}{\bf ---}) models are shown for all figures; for {\it E.coli} the gamma distribution (\textcolor{cyan}{\bf ---}), which performs better (measured by the Akaike weights) than either scale-free and the stretched exponential distributions.}
\caption{{\it Left panel}: The EBL background consisting of the GEBL and NIRBL7. {\it Right panel}: The \gray\opacity to a source at redshift$z$ = 0.122 due to $\gamma$-$\gamma$ interactions with the EBL. The shaded curves in the figure represent the contribution of the different wavelength regions (shown by the same shaded area in the left panel) to the total \gray\opacity. The dashed line represents the\gray\opacity caused by the GEBL alone (see Figs 1 and 2).}
\caption{{\it Left panel}: The observed \gray\spectrum of H1426+428 (Petry et al. 2002; filles diamonds) corrected for absorption by only the GEBL (open circles), the GEBL+NIRBL7 (filled squares), and the GEBL+NIRBL9 (open squares) realizations of the EBL.{\it Right panel}: The same as the left panel for PKS2155-304. The data point represent the H.E.S.S. observations by Aharonian et al. (2005)}
\caption{The observed \gray\spectrum of PKS2155-304 (Aharonian et al. 2005) corrected for the various realizations of the EBL (see Figure 6) is compared to the EGRET data of Vestrand et al. (1995, X symbols). A detailed discussion of the figure is in the text.}
\caption{ Non-Simultaneous OSSE, COMPTEL and EGRET \gray\spectra of 3C~454.3 (a) and CTA~102 (b). The solid lines represent the EGRET spectra inside the 1$\sigma$ error limits in slope. The OSSE results (McNaron-Brown et al. 1995) are from 4 VPs in 1994 for 3C~454.3, and from 3 VPs in 1994 for CTA~102 (see text). The EGRET spectra (Hartman et al. 1999) are from the first 4 CGRO Phases, i.e. time-averaged between April 1991 and November 1994 . }
\caption{\label{res4} $\triangle \left( \frac{T}{S} \right)$ as a function of sky fraction. The dashed line is computed based on the exact method for resolution 4, while the points with error bars are based on the Monte Carlo method for resolution 6 The continuous lines represent the theoretical scaling of $\triangle \left(\frac{T}{S} \right)$. The lower line corresponds to the idealized scaling $\triangle \left( \frac{T}{S} \right) \propto f_{sky}^{-1/2}$, while the upper line is $\triangle \left( \frac{T}{S} \right) \propto f_{sky}^{-2}$ and fits better the actual results for $f_{sky}>0.7$.}
\caption{\label{ground} $\triangle \left( \frac{T}{S} \right)$ for HEALPix resolution 7 for a map confined to Earth's southern hemisphere, declinations $\delta<-30^{\circ}$. The graph is obtained by ignoring the reionization peak for the $B$-modes and using the Monte Carlo method, with $1\sigma$ error bars shown on the graph. The continuous line is a cubic spline through the considered points.} \end{figure} \section{Discussion and conclusions} The main conclusion we get from these results is that the Fisher matrix element $F_{T/S,T/S}$ does not depend linearly on the sky percentage considered in the map. Instead, it depends more like $F_{T/S,T/S} \propto f_{sky}^4$. This is in some way expected as the harmonic decomposition of the sky is nonlocal and cutting off some parts of the sky affects the rest of it also especially for the large scale where the $B$-modes are more important. It is not an exact power law dependence, but as it can be seen from Fig.\ref{res4}, the power law could be a good approximation. We fitted the points from 10\% to 100\% in sky fraction to a power law. There is a disagreement between this dependence and the one from FIG.8 of Ref. \cite{2002PhRvD..65b3505L}, where $F_{T/S,T/S} \propto f_{sky}^{1.7}$. The method proposed in Ref. \cite{2002PhRvD..65b3505L} uses projected $E$ and $B$-modes, which triggers loss of information in the low multipoles. In the pixel space, this loss increases for small sky patches. To give an idea of the way low multipoles affect our analysis, in the case where the effect of the $B$-modes in the reionization peak was not considered (Fig.\ref{ground}), $F_{T/S,T/S} \propto f_{sky}^{1.2}$. We considered different pixelizations by changing the resolution of our HEALPix maps and we can see from Fig.\ref{res4} that this does not affect much the noise levels, leaving the results unchanged. The way we choose to make the cut is mildly important in evaluating the results. If instead of using a cut based on the galaxy dust maps, we take a straight symmetric cut, parallel to the galactic equator, we get higher $F_{T/S,T/S}$ values, but the change is modest (Fig.\ref{cut}). \begin{figure}[!h] \includegraphics[width=3.2in]{cut.eps} \caption{\label{cut} The difference in $\triangle \left(\frac{T}{S} \right)$ for two different types of cuts. The continuous line is obtained for the parallel equatorial cut. Both results are obtained through the exact method for HEALPix resolution 4.}
\caption{The measured effective preferential attachment (PA) rules for (a) the actor-movie network in the 1950s ($\circ$), 1960s ({\color{red}$\square$}), 1970s ({\color{green}$\Diamond$}), 1980s ({\color{blue}$\vartriangle$}), 1990s ({\color{brown}$\vartriangleleft$}) and 2000s ({\color{purple}$\triangledown$}), and for (b) the astrophysics, (c) the condensed matter physics and for (d) the high energy physics scientist--article networks during years 1998 ($\circ$), 1999 ({\color{red}$\square$}), 2000 ({\color{green}$\Diamond$}), 2001 ({\color{blue}$\vartriangle$}), 2002 ({\color{brown}$\vartriangleleft$}) and 2003 ({\color{purple}$\triangledown$}). The PA rules are well fitted by a power-law $T_k \sim k^\alpha$ with a cutoff and they appear to be independent of time. Numerically, we observe $\alpha \approx 0.65$ for the actor network, $\alpha \approx 0.6$ for the astrophysics network and $\alpha \approx 0.75$ for the other networks. The solid lines are guides to the eye.}
\caption{Fluctuation absolute magnitude vs.\colour}
\caption{Frequency as a function of applied field $h$. Comparison between fully nonlinear modes (direct numerical simulation of eq (\ref{eq:1})) and experiments \cite{Silva04}. Parameters are $I=10$ mA and fitted $\epsilon=0.5$. Inset: maximum frequency (\textasteriskcentered) and current ($\circ$) versus contact radius.}
\caption{\label{pic1}Vortex contribution to the ohmic losses in the film (A) as a function of the parameter $S_0=dw/\lambda_C^2(\omega=0)$, $\Omega=\omega/\omega_0$, and (B) as a function of the temperature for FC (\full) and ZFC (\dotted) regimes for the frequency $\omega/2\pi=1$~GHz. To avoid crowding, ZFC curves are multiplied by factor 4.}
\caption{Experimental results: the trap parameters ($a$) atom number and $b$) $1/ \sqrt{e}$ minor axis of the ellipse) are measured with a CCD camera looking from the (1, 1, 1)-direction, for 3 different single wave Rabi frequencies $\Omega=0.8\cdot \Gamma$~({\black \large +}), $\Omega= \Gamma$~({\blue \large +}), $\Omega=1.5\cdot \Gamma$~({\red \large +}). The peak density in the trap is $\sim 3 \cdot 10^{10}$ at. cm$^{-3}$.}
\caption[\RSi]{The automated construction used for defining \RSi. Green lines locate the \SiIIred and \SiIIblue \SiII lines. The blue lines locate the regions where the minima are searched for. The red diamonds locate the three maxima used to draw the two purple reference lines, and the two purple arrows represent $d_\textrm{red}$ and $d_\textrm{blue}$ used to compute $\textrm{\RSi}=\frac{d_\textrm{blue}}{d_\textrm{red}}$.}
\caption{Spectral energy distribution of the diffuse background radiation at $ z = 0$. Error bars show data points, triangles show lower limits from number counts and the inverted triangle is an upper limit from \gray observations (see text). The upper and lower solid lines show our fast evolution and baseline evolution predictions, and the dotted lines show our extensions into the optical--UV, as described by SS98. The steeply dropping solid line near $10^{12}$ Hz is the spectrum of the CMB.}
\caption{The optical depth of the universe from the IBL (fast evolution case) and the CMB as well as the total optical depth as a function of energy for a \gray\source at a redshift of 3. It can be seen that the contribution to the optical depth from the IBL dominates at lower\gray\energies and that from the CMB photons dominates at the higher energies. The dashed curve is for the IBL contribution alone and the dotted curve is for the CMB contribution alone.}
\caption{The optical depth of the universe to \grays from interactions with photons of the IBL and CMB 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 IBL cases and the dashed lines are for the baseline IBL cases.}
\caption{The optical depth of the universe to \grays from inteactions with photons as in Figure \protect \ref{taufam} but truncated at $\tau = 10^3$ to show detail.}
\caption{The optical depth of the universe, $\tau (E_{\gamma}, z)$, given as a continuous function of \gray energy and redshift for the fast evolution IBL case with interactions with the CMB included.}
\caption{The critical optical depth $\tau = 1$ as a function of \gray energy and redshift for the fast evolution (solid curve) and baseline (dashed curve) IBL cases. Areas to the right and above these curves correspond to the region where the universe is optically thick to \grays .}
\caption{The \gray data from the {\it H.E.S.S.} for PKS2155-304 compared with the theoretical spectrum for PKS2155-304 calculated by assuming an unabsorbed source spectrum proportional to $E^{-2}$ and multiplying by $e^{-\tau}$ using $\tau(z=0.117)$ for the fast evolution model based on Perez-Gonzalez \etal (2005) as discussed in the text.}
\caption{The unit square $\Omega_1$ and the isosceles triangle $\Omega_2$. Here and further on, \textcolor{red}{red solid lines \full} denote the Dirichlet boundary conditions and \textcolor{blue}{blue dashed lines \dashed} denote the Neumann ones.\label{fig:trivialex}}
\caption{Transmittance of VHE \grays\as a function of\gray\energy. Solid line:$(R, z, \alpha) = (0$ kpc, 0 kpc, $0^\circ)$ -- $L = 8.5$ kpc; Dashed line: $(R, z, \alpha) = (20$ kpc, 0 kpc, $90^\circ)$ -- $L = 21.8$ kpc; Dash-dotted line: $(R, z, \alpha) = (20$ kpc, 0 kpc, $180^\circ)$ -- $L = 28.5$ kpc. Thick lines give the total transmittance curve including the ISRF and CMB. Left-most thin lines give the transmittance for the ISRF only; right-most thin lines for the CMB only. \vspace{1\baselineskip} \label{fig3}}
\caption{The spectrum obtained from a single particle gyrating in a magnetic field from the code ({\it \grey}). The {\black} line shows the theoretical synchrotron spectrum. The dashed line indicates the power-law slope $2/3$. The inset in the upper left corner shows that the spectrum consists of a discrete set of spikes that are integer overtones of the gyrofrequency, in this case $\omega_B=0.094$ (arbitrary units) ({\it dotted vertical lines}). The particle has a Lorentz factor $\gamma=5$. All numbers are in simulations units.}
\caption{(a) Impurity concentration dependence of the superconducting transition temperature, determined from field-cooled magnetization measurements under $\mu_0H=5\times10^{-4}\mbox{T}$ ($\bigtriangleup$ and \Squaresteel\data) and from the electrical resistivity versus temperature curves ($\bigtriangledown$ data). Solid squares correspond to nonmagnetic (Lu) impurities and open symbols to magnetic (Pr) impurities. The solid line is a fit of the Abrikosov and Gor'kov approach.\cite{AG} (b)~An example, correponding to a La-1$\mbox{}\,\mbox{}$at.\%Pr alloy, of the as-measured magnetization (over $H$) versus temperature curves, showing directly the differences between the data and the background \MBTH\(solid line). These data were obtained under $\mu_0H=0.05\mbox{T}$, which is well inside the low-field regime for which \DM\is linear on the field amplitude. The temperature region shown corresponds well to the reduced temperatures covered in fig.~2(a). (c)~Magnetization over$H$ in the background temperature region for the same alloy.}
\caption{(a) Fluctuation-induced magnetization (over $H\xideo T$) versus reduced temperature in all the La-Pr alloys studied in this work and also for pure La and a La-1$\mbox{}\,\mbox{}$at.\/\%Lu alloy. These data were obtained under low magnetic fields, $\mu_0H=0.05\mbox{T}$. The covered $\epsilon$-range covered corresponds well to the accessible experimental window: closer to \Tc\the data are mainly affected by\Tc-inhomogeneities, whereas above $\varepsilon\simeq0.1$ they are mainly affected by background uncertainties. Outside this $\varepsilon$-region the experimental uncertainties could become even bigger than 100\%. The solid line is the GGL prediction, and the dashed lines the are GGL prediction plus a contribution proportional to the superfluid density and magnetic impurity concentration (see main text for details). (b)~Experimental fluctuation-induced magnetization, relative to the GGL predictions\cite{\citasSyS,Tinkham,MosqueiraPRL}, as a function of impurity content $x$ in pure-La \mbox{(\protect\circulo)}, La-Pr \mbox{($\protect\triangle$)} and La-Lu \mbox{(\protect\cuadrado)}, at $\varepsilon=0.06$ and $\mu_0H=0.05\mbox{T}$. The superconductors without magnetic impurities follow well the GGL approach (solid line), while in those with magnetic impurities $\left|\DM\right|$ increases, in essence linearly, with $x$. The dashed and dot-dashed lines are explained in the main text. } \end{figure} } %%%%%%---tabla---%%%%%%%% \newcommand{\tablauno}{ \begin{table} \begin{center}\mbox{}\vskip-0.8cm\mbox{}\\ \begin{tabular}{lccccccc} \hline \hline Sample & \Tc & $\Delta\Tc$ & $\mu_0H_{c2}(0)$ & $\xi(0)$ & $\kappa(\Tc)$&$\ell$&$d_{imp}$ \\ &(K) &(K) &(T) & (\AA) & &(\AA)&(\AA)\\ \hline La & 5.85 & 0.16 & 0.8 & 200 & 4.4 & 265 & \\ La-0.5 at.\% Pr & 5.69 & 0.12 & 0.9 & 190 & 5.6 & 150 & 19 \\ La-1 at.\% Pr & 5.51 & 0.25 & 0.9 & 180 & 6.0 & 110 & 15\\ La-2 at.\% Pr & 5.40 & 0.18 & 1.0 & 180 & 6.4 & 75 & 12\\ La-2 at.\% Lu & 5.88 & 0.26 & 0.9 & 190 & 4.2 & 210 & 12\\ \hline \hline \end{tabular} \end{center} \caption{\label{tabla} Summary of basic parameters of the La-Pr superconductors studied in this work, obtained from magnetization and electrical resistivity measurements.\cite{tobepublished} For comparison, we included also one of the pure La samples and one of the La-Lu alloys. The average distance between impurities $d_{imp}$ was estimated from the nominal sample compositions. $\ell$ is the mean free path of the normal electrons extrapolated to $T=0 {\rm K}$, and $\kappa(\Tc)$ the Ginzburg-Landau parameter at the critical temperature. The uncertainties in $\xi(0)$, $\kappa(\Tc)$ and $\ell$ are of about $20\%$.} \end{table} } %%%%--preamble with title, authors etc.: \newcommand{\titulo}{ Observation of enhanced fluctuation diamagnetism\\ in lanthanum superconductors with\\ dilute magnetic impurities } \newcommand{\autores}{F\'elix~Soto, Luc\'{\i}a~Cabo, Jes\'us~Mosqueira,\\ Manuel~V.~Ramallo, Jos\'e~A.~Veira and F\'elix~Vidal} \newcommand{\direccion}{ Laboratorio de Baixas Temperaturas e Superconductividade,\footnote{Unidad Asociada al Instituto de Ciencia de Materiales de Madrid, CSIC, Spain} Departamento de F\'{\i}sica da Materia Condensada,\\ Universidade de Santiago de Compostela, E15782 Spain. } \begin{center} \Large\bf \titulo\\ \end{center}\mbox{}\vspace{-1cm}\\ \begin{center}\large\autores\end{center} \begin{center}\large\it\direccion\end{center} %%%%---abstract--- \mbox{}\vskip2cm{\bf Abstract. } The fluctuation-induced diamagnetism \DM, associated with the presence of precursor Cooper pairs in the normal state, has been measured in lanthanum with dilute magnetic (Pr) and nonmagnetic (Lu) impurities. It is found that while for pure La and La-Lu alloys \DM\ agrees, as expected, with the theoretical predictions, it is much larger for La-Pr alloys (around a factor 5 for La-2$\mbox{}\,\mbox{}$at.\%Pr). These results suggest the existence of an indirect contribution to \DM\ arising from the interaction between fluctuating Cooper pairs and magnetic impurities. \newpage \setlength{\baselineskip}{18pt} %%%%---main body of the article The magnetic susceptibility measurements of \Matias\ and coworkers\cite{Matias} showing the decrease of the superconducting transition temperature of lanthanum with dilute magnetic rare earth impurities, and the explanation of these effects by Abrikosov and Gor'kov\cite{AG} in terms of pair-breaking, opened about 45 years ago one of the still at present most interesting and studied issues of correlated electron systems: the interplay between magnetism and superconductivity.\cite{Fisher,Osborne} A natural question directly related to these pioneering results but to our knowledge not yet addressed until now is how the magnetic impurities will affect the precursor diamagnetism associated with the presence of fluctuating Cooper pairs created above any superconducting transition by the unavoidable thermal agitation. In fact, this fluctuation-induced diamagnetism above the superconducting transition may be used as an unique probe to study the competition between pairing correlations and magnetic order without entering into the fully-superconducting state, which would hide the response of the magnetic ions to external fields. We present in this Letter measurements of the fluctuation-induced diamagnetism, $\DM\equiv M(T)-M_B(T)$, in La with dilute (up to 2$\mbox{}\,\mbox{}$at.\%) magnetic (Pr) and nonmagnetic (Lu) impurities. Here $M(T)$ and $M_B(T)$ are the measured and, respectively, the background or bare magnetizations. It is found that while for pure La and La-Lu alloys \DM\ agrees with the calculations on the grounds of the Gaussian-Ginzburg-Landau (GGL) approach for isotropic 3D superconductors\cite{\citasSyS,Tinkham,Gollub,MosqueiraPRL}, the \DM\ amplitude is much larger in the La-Pr alloys (about 500\% for impurity concentrations of 2$\mbox{}\,\mbox{}$at.\%). The results for La-Lu alloys just confirm previous experiments in other dirty low-\Tc\ superconductors\cite{Tinkham,Gollub,MosqueiraPRL} and they may be easily understood by just assuming that the fluctuating Cooper pairs in the normal state are also protected, as the Cooper pairs below the superconducting critical temperature \Tc, by the Anderson theorem for symmetric perturbations.\cite{Anderson} These results in La-Lu superconductors provide then a crucial check of the reliability of our measurements of \DM\ in the lanthanum-based alloys. In contrast with these conventional results in La-Lu alloys, the \DM-enhancement in presence of magnetic ions suggests not only that the fluctuating Cooper pairs in conventional (singlet $s$-wave pairing) BCS superconductors are very robust to antisymmetric perturbations but also the existence of an indirect contribution to \DM\ arising from the interaction between Cooper pairs and magnetic impurities. These results may have implications in other scenarios where superconducting fluctuations and magnetic order coexist\cite{Fisher,Osborne}, in particular when the superconductivity is magnetically mediated\cite{\unconventional}, perhaps including the high-\Tc\ cuprates\cite{Osborne}. In addition to its interest as a natural extension of the seminal experiments of \Matias\ and coworkers,\cite{Matias} our choice of the magnetization to study the interplay between fluctuating Cooper pairs in the normal state and magnetic impurities was motivated by the fact that, as known since the early experiments of Tinkham and coworkers in low-\Tc\ superconductors without magnetic impurities,\cite{Tinkham,Gollub} the fluctuation-induced magnetization probably is the best observable to probe the superconducting fluctuations in any isotropic bulk (3D) superconductor. In contrast with most of the effects of the superconducting fluctuations on other observables, $\left|\Delta M\right|$ increases not only with the density of fluctuating Cooper pairs but also with their size, \ie, with the superconducting coherence length amplitude (extrapolated to $T=0\mbox{K}$), \xideo. As a consequence, $\left|\Delta M/M_B\right|$ may take relatively high values at easily accessible temperature-distances from the transition in isotropic low-\Tc\ superconductors, even bigger than in high-\Tc\ superconductors, due to their much larger \xideo.\cite{MosqueiraPRL} In fact, for the superconductors studied in the present Letter we will obtain values of $\left|\Delta M/M_B\right|$ well comparable (about one order of magnitude smaller in the worst case) with those of optimally-doped \mbox{YBa$_2$Cu$_3$O$_{7-\delta}$}.\cite{medidasYBCO} These experimental advantages appear to be crucial when compared with the difficulties that may arise if one uses other observables, very in particular the electrical conductivity, to study the interplay between superconducting fluctuations and magnetic impurities in low-\Tc\superconductors: mainly in bulk materials, the ratio between the corresponding paraconductivity,\Ds, and the nonfluctuating (background) conductivity, $\sigma_B$, may be orders of magnitude smaller than $\left|\DM/M_B\right|$. In addition, $\sigma_B$ may be much more affected than $M_B$ by stoichiometric and structural inhomogeneities. These difficulties probably affect deeply the few (to our knowledge) attempts published until now to study through the electrical conductivity the interplay between superconducting fluctuations and magnetic impurities in low-\Tc\ superconductors.\cite{Spahn,Rapp} The fitness of the La alloys to measure \DM\ in presence of magnetic impurities is mainly due to the fact that these compounds still have a quite large \xideo, of the order of 200\AA. Complementarily, for some magnetic impurities (very in particular Pr), their normal state magnetization is expected to remain relatively moderate, even under impurity concentrations $x$ as important as 2$\mbox{}\,\mbox{}$at.\%, the larger impurity concentrations used by \Matias\ and coworkers in their measurements of $\Tc(x)$\cite{Matias}. In addition, at present high-quality La alloys are commercially available. The polycrystalline \mbox{La$_{100-x}$Pr$_x$} and \mbox{La$_{100-x}$Lu$_x$} alloys studied in this Letter, with $0\leq x\leq 2$, were supplied by Goodfellow and Alfa Aesar, and their impurity concentration was controlled to better than 0.1\%. The superconducting transition width, \DTc, as determined from electrical resistivity and field-cooled magnetic susceptibility measurements remains below $\sim$0.25\/K for all the samples, which confirms their good stoichiometric and structural quality. These measurements, as well as those performed to determine other general characteristics of the samples (very in particular \Hcdoso\ and then \xideo), are detailed elsewhere\cite{tobepublished}. Here we present in \tabla\ some of the main sample parameters and in fig.~1(a) the impurity concentration dependence of \Tc, which for the La-Pr alloys provides a direct measure of the pair-breaking effects induced by the magnetic impurities. These \Tcx\ data are in agreement with both the results of \Matias\ and coworkers\cite{Matias} and the Abrikosov-Gor'kov approach\cite{AG}. The solid line in fig.~1(a) is a fit of this approach, which at low concentrations reduces to $\Tc(x=0)-\Tc(x)=\hbar/\kB\tau_\phi(x)$, where $\hbar$ is the reduced Planck constant, \kB\ is the Boltzmann constant and $\tau_\phi(x)=\tau_\phi(x=1)/x$ is the phase pair-breaking time. This leads to $\tau_\phi(x=1)\simeq3.4\times10^{-11}\mbox{s}$, which is much larger than the relaxation time of the normal electrons deduced from resistivity measurements\cite{tobepublished} ($\tau\sim10^{-14}\mbox{s}$). \Tcx\ for the alloys with nonmagnetic Lu-impurities is almost constant, the Cooper pairs in these last alloys being protected, as is well known, by the Anderson theorem.\cite{Anderson} An example of the temperature behaviour above \Tc\ of the measured and of the background magnetizations is presented in figs.~1(b) and (c). These measurements were performed under constant magnetic field amplitude (and in the low-field regime, $H\ll\Hcdoso$) with a commercial, SQUID based, magnetometer (Quantum Design's MPMS) and by using cylindrical samples of diameter and height $\sim5\mbox{mm}$, the maximum volume allowed by our magnetometer. The demagnetizing effects are negligible above \Tc. Other experimental details are similar to those in magnetization measurements in other low-\Tc\ superconductors.\cite{MosqueiraPRL} The solid line in figs.~1(b) and (c) is the background or bare magnetization (over $H$), obtained by fitting in the temperature region bounded by $1.5\Tc\leq T\leq3\Tc$ the function $\MBTH=\chi_0+BT+C/T$ (with $\chi_0$, $B$ and $C$ as free parameters). In this region the agreement between \MBTH\ (solid line) and the data is excellent, the maximum deviation being below 0.05\%. Let us stress that even above this background fitting region, up to $5\Tc$, the agreement with the data still remains quite good, the maximum deviation being below 5\%. The above background functionality is expected to be a good approximation for (anti)ferromagnetic materials at temperatures well above the Curie or N\'eel temperatures \Tmag,\cite{libromagnetismo} which in superconductors with diluted magnetic impurities may be estimated through the relationship\cite{deGennes} $\Tmag(x)\sim\Tc(x=0)-\Tc(x)$. In our samples this leads to $\Tmag\lsim0.5\mbox{K}\ll\Tc$. From curves similar to those presented in fig.~1(b), we obtain the fluctuation-induced magnetization versus reduced-temperature curves, \DMe, where $\varepsilon\equiv\ln(T/\Tc)$. In fig.~2(a) we present some examples of these \DMe\ curves, very in particular those that correspond to the La-Pr alloys. These data are well inside the $\varepsilon$-range where the GGL approach is expected to be applicable.\cite{\citasSyS,Tinkham,Gollub,MosqueiraPRL,medidasYBCO} The solid line corresponds to the GGL calculations for isotropic 3D superconductors without magnetic impurities in the zero-field limit,\cite{\citasSyS,Tinkham,MosqueiraPRL} %%%%%%%%%%% \begin{equation} \frac{\DMSe}{H\xideo T} = \frac{-\mu_0\kB}{3\phi_0^2} \left[ \frac{{\rm arctan}\sqrt{\frac{\mbox{$\varepsilon^C-\varepsilon$}}{\mbox{$\varepsilon$}}}} {\sqrt{\varepsilon}} - \frac{{\rm arctan}\sqrt{\frac{\mbox{$\varepsilon^C-\varepsilon$}}{\mbox{$\varepsilon^C$}}}} {\sqrt{\varepsilon^C}} \right], \label{DMSS} \end{equation} %%%%%%%%%%% where $\eC\simeq0.6$ is the BCS value of the cutoff constant\cite{EPLVidal} and $\phi_0$ is the flux quantum. As visible in this figure, the data in pure La and in La-Lu alloys agree with \eq{DMSS} well within the experimental uncertainties, which are of the order or below 40\%\cite{cuarenta}, an agreement that was already observed in other low-\Tc\ superconductors with nonmagnetic impurities.\cite{MosqueiraPRL} As stressed before, this result may be easily understood by just assuming the applicability also above \Tc\ of the immunity predicted by Anderson\cite{Anderson} of the Cooper pairs to symmetric perturbations. So, these last results also show the reliability of our present experiments to probe the superconducting fluctuations above \Tc\ in La alloys. The results of fig.~2(a) show also that in presence of magnetic impurities {\it both} the {\it amplitude} and the {\it $\varepsilon$-dependence} of $\DMe/H\xideo T$ differ from the GGL calculations\cite{\citasSyS,Tinkham,MosqueiraPRL}, even for the alloy with the lower Pr concentration: The $\DMe/H\xideo T$ amplitudes remain much larger, about five times for the La-2~at.\%Pr alloy, and also their reduced-temperature dependence is smoother. As illustrated quantitatively in fig.~2(b), this enhancement is essentially proportional to the magnetic impurity concentration, in striking contrast with the well established and well understood decrease of the superconducting transition temperature when such a concentration increases\cite{Matias,AG,Fisher} (see fig.~1(a)). In analyzing the experimental findings summarized in figs.~2(a) and (b) we must start by checking if such an strong \DM-enhancement is just an effect due to \Tc-inhomogeneities at long-length scales (bigger than \xideo), in turn associated with stoichiometric and structural defects.\cite{Vidalimpurezas} However, this extrinsic effect is ruled out not only by the high quality of the samples but mainly by the fact that the La-Lu alloys display magnetic and electrical resistivity transition widths, \DTc, similar to those of their La-Pr counterparts (see \tabla). Also, a rounding of the magnetic transition due to local \Tc-inhomogeneities, induced by the magnetic ions (inhomogeneities either isolated\cite{\LOref} or Josephson tunneling-connected\cite{\Kresinref}), may be discarded because in our La-Pr alloys the mean distances among impurities, \di, are in the range $12-19$\AA, which is well below the \xideo\ values $\sim200$\AA\ (see \tabla). The amplification effects associated with a reduction of the superconducting fluctuations' dimensionality (in turn due to a confinement of these fluctuations between the impurity interspaces) may be also discarded. This is illustrated in fig.~2(b), where the dot-dashed line corresponds to the fluctuation magnetization that may be obtained by supposing that the magnetic impurities would allow only 0D superconducting fluctuations occurring in the interspace between magnetic impurities. For such a \DMe\ in 0D we used\cite{Tinkham} %%%%%%%%%%%%%%%%%%% $ \DMZD(\epsilon)=(\xideo/\di\varepsilon^{1/2})\DMSe. $ %%%%%%%%%%%%%%%%%%%% Another possible amplification effect, opposite in origin to a superconducting fluctuations' confinement, would be that these fluctuations embrace part of the magnetic ions, cancelling their contribution to the magnetization. A crude estimation of these effects may be done by just identifying $\left|\DMS/H\right|$ with the effective volume fraction occupied by the evanescent superconducting ``droplets''. The resulting contribution to the magnetization is then $\MB\DMS/H$ which, taking into account that in our samples $10^{-4}\lsim\MB/H\lsim10^{-3}$ near the transition, leads to a negligible effect. Finally, we also note that a mechanism of internal-field screening different to Meissner-like currents, which involves spin alignment of the quasiparticles composing the Cooper pair, has been recently proposed, for $T<\Tc$ and noninteracting magnetic impurities.\cite{EPL} However, when applied to $T>\Tc$ the resulting change in magnetization is again about three orders of magnitude smaller than our observations. In absence of a theoretical approach for the superconducting fluctuations in presence of magnetic impurities, and as a {\it first crude} attempt to establish {\it empirically} the physical origin of the observed \DM-enhancement, we have checked if it is possible to separate the measured \DMe\in two contributions: the direct GGL\SySnombre\term and an indirect contribution proportional to both the density of fluctuating Cooper pairs and the magnetic impurity concentration. This check has been summarized in figs.~2(a) and 2(b), where the dashed lines were obtained using the empirical expression%%%%%%% \begin{equation} \DMe=\DMSe+\Aind \, H\, x\,\nse, \label{empirico} \end{equation} %%%%%%%% where \Aind\is a constant and\nse\is the superfluid density (in dimensionless units) for isotropic 3D superconductors in the zero-field limit,\cite{EPLVidal} %%%%%%% \begin{equation} \nse \hskip-0.2em=\hskip-0.2em \frac{\kB \mu_0 {\rm e}^2 T\xideo}{\hbar^{2}\sqrt{\varepsilon}} \left[\sqrt{\frac{\varepsilon^C-\varepsilon}{\varepsilon}} - {\rm arctan}\sqrt{\frac{\varepsilon^C-\varepsilon}{\varepsilon}} \right]\hskip-0.2em. \label{eqns} \end{equation} %%%%%%%% The dashed lines in fig.~2(a) correspond to fits to the La-Pr alloy data of the above $\DMe$ expression, with \Aind\as the only free parameter (we constrained the rest of parameters to be compatible with\tabla, and also used $\varepsilon^C=0.6$ for all the fits). As it may be seen in this figure, the agreement between the experimental results and \eq{empirico} is excellent, for {\it both} the amplitude {\it and} the $\varepsilon$-behaviour, and this using the same value $\Aind\simeq-0.75$ for all the samples. The above results lead to two complementary conclusions: First, the presence of the direct GGL \SySnombre\contribution in the La-Pr alloys suggests the robustness of the fluctuating Cooper pairs to dilute magnetic impurities. Although unexpected when compared to the decrease of\Tcx\shown in fig.~1(a), this behaviour agrees with recent\DM-measurements in Pb-In alloys under high reduced-fields,\cite{Soto} which also suggests the robustness of the fluctuating Cooper pairs to the antisymmetric nature of an applied magnetic field. Complementarily, the presence of an indirect contribution to \DMe\in the La-Pr alloys suggests that the fluctuating Cooper pairs modify the coupling between the magnetic ions. In the case of dilute alloys such a coupling is mediated by the electronic sea (RKKY model).\cite{libromagnetismo,RKKY} It may be then expected that the changes induced in the electrons' spin susceptibility by the Cooper pair formation will also affect the magnetization due to the magnetic ions. In fact, such an effect was early proposed below \Tc\\cite{GRFF}, and the corresponding change in the magnetization above \Tc\may be expected to be proportional to both the superfluid density and the concentration of magnetic ions, in agreement with our experimental findings. These results suggest that the energetic balance between the magnetic and the superconducting orders could be shifted by the presence of precursor Cooper pairs. Although these crude ideas need to be confirmed both experimentally and theoretically, the experimental findings presented in this Letter already stress that the superconducting fluctuations may play an unexpected important role in the interplay between magnetism and superconductivity, mainly if the latter is magnetically mediated.\cite{Osborne,Saxena} \mbox{} {\bf Acknowledgments} We acknowledge useful conversations and correspondence with J.B.~Goodenough and K.~Maki on this topic. This work was supported by the CICYT, Spain (grant no.\MAT2004-04364), the Xunta de Galicia (PGIDT01PXI20609PR), and Uni\'on Fenosa (220/ 0085-2002). LC acknowledges financial support from Spain's Ministerio de Educaci\'on y Ciencia through a FPU grant.\mbox{} \begin{thebibliography}{99} \nonfrenchspacing \newcommand{\cursiva}{\it} \newcommand{\negrita}{\bf} \bibitem{Matias}\autor{B.T.}{\Matias}, \autor{H.}{Suhl}\y \autor{E.}{Corenzwit}, \revista{Phys. Rev. Lett.}{1}{92}{94}{1958}; \autor{R.A.}{Hein} \etal, \revista{{\it ibid.}}{2}{500}{502}{1959}. \bibitem{AG}\autor{A.A.}{Abrikosov}\y \autor{L.P.}{Gor'kov}, \revista{Sov. Phys. JETP}{12}{1243}{1253}{1961}. \bibitem{Fisher}See, \eg, \autor{\O.}{Fisher}, in \librocorto{Ferromagnetic Materials, Vol.~5, {\rm edited by \editor{K.H.J.}{Buschow}\y \editor{E.P.}{Wohlfarth} }}{Elsevier}{Amsterdam}{1990}, and references therein. \bibitem{Osborne}See, \eg, {\it Correlated Electron Systems} (\revista{Science}{288}{pp.461-482}{}{2000}). \bibitem{SSuno}\autor{H.}{Schmidt}, \revista{Z. Phys.}{216}{336}{345}{1968}; \autor{A.}{Schmid}, \revista{Phys. Rev.}{180}{527}{529}{1969}. \bibitem{Tinkham} For a review see, \eg, \autor{J.}{Skocpol} \y \autor{M.}{Tinkham}, \revistacorto{Repts. Prog. Phys.}{38}{1049}{1975}; \autor{M.}{Tinkham}, \libro{Introduction to superconductivity}{\chap~8}{McGraw-Hill}{N.Y.}{1996}. \bibitem{Gollub} \autor{J.P.}{Gollub}, \autor{M.R.}{Beasley}, \autor{R.S.}{Newbower}\y \autor{M.}{Tinkham}, \revista{Phys. Rev. Lett.}{22}{1288}{1291}{1969}; \autor{J.P.}{Gollub}, \autor{M.R.}{Beasley}\y \autor{M.}{Tinkham}, \revista{{\it ibid.}}{25}{1646}{1649}{1970}; \autor{J.P.}{Gollub} \etal, \revista{Phys. Rev. B}{7}{3039}{3058}{1973}. \bibitem{MosqueiraPRL}\autor{J.}{Mosqueira}, \autor{C.}{Carballeira}\y \autor{F.}{Vidal}, \revistacorto{Phys. Rev. Lett.}{87}{167009}{2001}. \bibitem{medidasYBCO}See, \eg, \autor{F.}{Vidal}\y \autor{M.V.}{Ramallo}, in \libro{The gap symmetry and fluctuations in high-\Tc\superconductors,{\rm edited by \editor{J.}{Bok} \etal}}{p.~443}{Plenum}{N.Y.}{1998}, and references therein. \bibitem{Anderson}\autor{P.W.}{Anderson}, \revista{J. Phys. Chem. Solids}{11}{26}{28}{1959}. \bibitem{Saxena} See, \eg, \autor{S.S.}{Saxena} \etal, \revista{Nature}{406}{587}{592}{2000}. \bibitem{Spahn}\autor{E.}{Spahn}\y \autor{K.}{Keck}, \revista{Physica B}{165-166}{1357}{1358}{1990}. The samples measured in this work are Al films with a magnetic (Ni) covering, and therefore probably are highly inhomogeneous. \bibitem{Rapp} \autor{P.}{Lindqvist}, \autor{A.}{Nordstr\"om}\y \autor{\"O.}{Rapp}, \revista{Phys. Rev. Lett.}{64}{2941}{2944}{1990}. As shown in Ref.\cite{Vidalimpurezas}, the anomalies reported here may be easily explained just in terms of extrinsic \Tc-inhomogeneities at long length scales (much larger than \xideo) and non-uniformly distributed. \bibitem{tobepublished}{\autor{F.}{Soto} \etal, J. Phys. Chem. Solids (in press). An extended version of this paper may be found in cond-mat/0511033.} \bibitem{libromagnetismo} See, \eg, \autor{J.}{Crangle}, \libro{Solid state magnetism}{\chaps~6 and~7}{Edward Arnold}{London}{1991}. The first two $M_B/T$ contributions, $\chi_0$ and the small $BT$, are usual in materials without magnetic impurities (and were also used in the analyses of \DM\in low-\Tc\superconductors without magnetic impurities as,\eg, Pb-In alloys\cite{MosqueiraPRL}). The magnetic impurities, at the low magnetic fields used in this work (well smaller than the saturation fields), will produce additional $M_B/T$ contributions $\propto(T\pm\Tmag)^{-1}$, which in our $\Tmag\ll T$ case may be safely approximated as $C/T$. (Note that the correlation of very diluted magnetic impurities will be mediated by the electronic sea and hence may be ferro- or antiferromagnetic, see Ref.\cite{RKKY}). We obtained for all our La-Pr alloys $C$ values in good agreement with those corresponding to their nominal concentration. \bibitem{deGennes} See, \eg, \autor{P.G.}{de Gennes}, \libro{Superconductivity of metals and alloys}{p.~265}{Benjamin}{N.Y.}{1966}. \bibitem{EPLVidal} \autor{F.}{Vidal} \etal, \revista{Europhys. Lett.}{59}{754}{760}{2002}. Let us stress here that as the reduced-temperature range studied in this work is relatively close to \Tc\($\varepsilon\lsim0.1$) the use of a conventional momentum cutoff\cite{Tinkham} would not appreciably modify the theoretical \DM. \bibitem{cuarenta}These uncertainties are mainly due to the background extraction. They have been estimated by varying by 10\% the temperature location of the background fitting region. \bibitem{Vidalimpurezas}{\autor{F.}{Vidal} \etal, in \libro{Materials science, fundamental properties and future electronic applications of high-\Tc\superconductors,{\rm edited by \editor{S.L.}{Drechsler}\y \editor{T.M.}{Mishonov}}}{p.~289}{Kluwer}{Dordrecht}{2001}; a recently updated version may be found in cond-mat/0510467.} \bibitem{LOuno}{\autor{A.}{Larkin}\y \autor{Yu.N.}{Ovchinnikov}, \revista{Sov. Phys. JETP}{34}{651}{655}{1972}}. \bibitem{Kresin}{\autor{Yu.N.}{Ovchinnikov}, \autor{S.A.}{Wolf}\y \autor{V.Z.}{Kresin}, \revista{Phys. Rev. B}{60}{4329}{4333}{1999}. } \bibitem{EPL}{\autor{F.S.}{Bergeret}, \autor{A.F.}{Volkov}\y \autor{K.B.}{Efetov}, \revista{Europhys Lett.}{66}{111}{117}{2004}.} \bibitem{Soto}{\autor{F.}{Soto} \etal, \revistacorto{Phys. Rev. B}{70}{060501}{2004}. \bibitem{GRFF}{\autor{L.P.}{Gor'kov}\y \autor{A.I.}{Rusinov}, \revista{Sov. Phys. JETP}{19}{922}{931}{1964}; \autor{P.}{Fulde}\y \autor{R.}{Ferrel}, \revista{Phys. Rev.}{135}{A550}{A563}{1964}.} \bibitem{RKKY}\autor{M.A.}{Ruderman}\y \autor{C.}{Kittel}, \revistacorto{Phys. Rev.}{96}{99}{1954}; \autor{T.}{Kasuya}, \revistacorto{Progr. Teor. Phys.}{16}{45}{1956}; \autor{K.}{Yosida}, \revistacorto{Phys. Rev.}{106}{893}{1957}. } \end{thebibliography} \referee{ \newpage\mbox{}\\ \figurauno \mbox{}\\ \newpage\mbox{}\\ \figurados \mbox{}\\ \newpage\mbox{}\\ \tablauno\mbox{}\\ } \end{document} }
\caption{Speed of sound as a function of $\tau$. (\lbullet) and (\ldiamond) are results for $k=(1,0)$ and $(0,1)$, respectively. The solid line is the theoretical prediction given by Eq. (\ref{sound}). The inset shows the dynamical structure factor as a function of $\omega$ for $k=(2,0)$, $\tau=0.2$. Parameters: $L/a=128$, $M=5$, $k_BT=1.0$.}
\caption{$S(\overline{k},t=0)/\rho$ as a function of $\tau$ (\lbullet). (\ldiamond) are results obtained by numerically evaluating the derivative of the pressure measured using the microscopic stress tensor. The solid line is a plot of Eq. (\ref{structure}). $\overline{k}$ is the lowest wave vector.}
\caption{\label{ealphadriven} Plot of $\sigma^2_E$ versus the restitution coefficient $\alpha$ for $N=100$ ($\bigcirc$) and $N=1000$ (\textcolor{red}{$\square$}) particles driven by the stochastic thermostat. The result of the calculation assuming uncorrelated velocities (\ref{sigmauncorr}) is plotted in dashed line.}
\caption{\label{ealphacooling} Plot of $\sigma^2_E$ versus the restitution coefficient $\alpha$ for $N=100$ ($\bigcirc$) and $N=1000$ (\textcolor{blue}{$\square$}) particles, for the Gaussian thermostat. The solid line shows the theoretical predictions of \cite{brey04a}, given by Eq.~(\ref{eq:sigmaegauss}).}
\caption{\label{figenpdfcool} pdf of the rescaled energy $\widetilde E$ when the restitution coefficient is $\alpha = 0.5$. The system is driven with the Gaussian thermostat with 100 ($\bigcirc$) and 1000 (\textcolor{blue}{$\square$}) particles. The solid line shows a rescaled gamma distribution with zero-mean, a variance equal to 1, and with a number of degrees of freedom $N_f$ given by $2 N /\sigma_E^2$. The dashed line is the Gaussian with zero-mean and unit variance.}
\caption{$j(t,0)$ (\full) of $v_n$ and the corresponding half space probability (\textcolor{red}{\broken}) as functions of time.}
\caption{Flow lines of the velocity field $j/\rho$ of $v_n$. The flow lines between \textcolor{red}{\broken} pass through $x=0$ in the negative direction. Two consecutive flow lines are separated by a probability of approximately $2.4 \cdot 10^{-3}$.}
\caption{\label{lc1}Multi-wavelength data from December, 2002. The \gray data points show per-night averages, in Crab units. Starred data points signify data taken from the HEGRA CT1 telescope, while crosses denote Whipple 10m data. The error bars on the RXTE PCA data are not shown as they are smaller than the symbol size, and have units of $10^{-10}$ erg cm$^{-2}$ s$^{-1}$ at 10 keV. The circled X-ray and \gray data points overlapped or were taken less than 5 min apart. The TeV \gray and RXTE photon indices show $\Gamma$, where $dN/dE \propto E^{-\Gamma}$. the ASM data are given in mCrab. In the optical band, open circles show the WIYN V band data, crosses show the Boltwood R band data, and 'x' denotes La Palma R band data. All of the optical data are in relative magnitude units. The open (filled) circles in the radio band show measurements that overlapped or were taken within 5 min of a TeV \gray observation (X-ray and TeV \gray observation). The radio data are given in Janskys.}
\caption{\label{dcf}Discrete correlation function of the complete X-ray and \gray data set. A positive time lag means the \gray flux precedes the X-ray flux.}
\caption{\label{xgamma}Plot of the TeV \gray versus X-ray flux correlation for measurements for all overlapping data sets.}
\caption{\label{nufnu}Mrk 421 Spectral Energy Distributions measured during the campaign. The data points show the radio to \gray data. The radio and optical spectra show the average fluxes observed during the campaign. At X-rays we show three spectra, one from the {\it RXTE} pointing with the highest flux observed during the campaign, one from the pointing with the lowest flux, and one spectrum at intermediate flux levels. The intermediate X-ray spectrum and the gamma-ray spectrum were determined using for both only the data taken during nights with simultaneous X-ray and gamma-ray observations. The solid curved lines show, for comparison, a low-flux and a high-flux energy spectrum measured with {\it BeppoSAX} (``Beppo'' Satellite per Astronomia X) during the 1998 flaring period \citep{Foss:00}. the long dashed lines show the results from a simple Synchrotron Self-Compton model with $\delta = 1000$, while the short dashed lines show the results with $\delta = 50$. }
\caption{\label{parameters}Parameters used for the TeV \gray energy estimator.}
\caption{(Color online) The four measured level-crossing fields\cite{fe12mag} (dashed vertical lines) are compared with the theoretical $dM/dH$ that result from $N=12$, $s=5/2$ Heisenberg rings with $k_B T / J = 0.01$. The theoretical data are shown for the following three choices of $J$ and $g$: $J = 31.9$ K and $g = 2.00$ ({\color{blue} $\diamond$}) from Ref.~\onlinecite{fe12chem}; $J = 40.7$ K and $g = 2.02$ ({\color{red} $\Box$}) from Ref.~\onlinecite{fe12mag}; $J = 35.2$ K and $g = 2.0$ ($\circ$) are our best estimates.}
\caption{\label{fig:angcor}(Color online) Calculated angular distribution, $W$, as a function of the angle $\theta_{\gamma_1-\gamma_2}$ between the two \gamrs\emitted in the\gamgam\cascade. The solid curve is the angular correlation for all$0^+ \rightarrow 2^+ \rightarrow 0^+$ decay sequences. The dashed (dotted) curve is the correlation between the two \gamrs\emitted from the$2_2^+ \rightarrow 2_1^+ \rightarrow 0^+$ sequence of \nuc{100}{Ru} (\nuc{102}{Ru}).}
\caption{\label{fig:datmo}(Color online) The \gray\spectra in coincidence with$590.8\pm2.5$ keV (a) and with $539.5\pm2.5$ keV (b). The 22 observed 591 - 540 keV events (shaded) from the \bb$(0^+\rightarrow0_1^+)$ decay of \nuc{100}{Mo} were obtained in 455 days of measuring time.}
\caption{\label{fig:datmo-5}(Color online) The \nuc{100}{Mo} \gamgam\coincidence data, compressed to include 5 keV per bin. Shown are the\gray\spectra in coincidence with$590.8\pm2.5$ keV (a) and with $539.5\pm2.5$ keV (b), with some of the prominent background lines identified (CS = Compton Scattering). Note the 22 observed 591 - 540 keV coincidence events from the \bb ($0^+\rightarrow0_1^+$) transition.}
\caption{\label{fig:moschemefull}Decay scheme for the \bb\decay of\nuc{100}{Mo} to higher excited states of \nuc{100}{Ru}, excluding the $0_1^+$ state. Shown are the subsequent gamma cascades and the corresponding \gray\energies and branching ratios.}