\caption{ Directly-observed (solid), and modified (dots) proton and electron spectra. The modified spectra are deduced from fits to antiproton and \gray data \cite{2004ApJ...613..962S}. % SMR2004 LIS marks local interstellar spectra. Data shown are from AMS01, BESS, CAPRICE94, IMAX, LEAP, HEAT, MASS91, SANRIKU. }
\caption{\gray spectrum of inner Galaxy ($330^\circ<l<30^\circ, |b|<5^\circ$) for model based on the directly-observed CR spectra and modified spectra shown in Fig.~\ref{protons_electrons_SMR2004}. Red bars: EGRET data, including points above 10 GeV, see \cite{2005A&A...444..495S}. %Strong etal Green bars: COMPTEL. Light blue line: bremsstrahlung, green line: inverse Compton scattering, red line: $\pi^0$-decay, black line: extragalactic background, dark blue line: total. This is an update of the spectra shown in \cite{2004ApJ...613..962S}. % SMR2004 }
\caption{Possible CR source distributions as function of Galactocentric radius $R$: pulsars (solid line), SNR (vertical bars), from \grays assuming constant H$_2$-to-CO relation (dotted line) \cite{2004A&A...422L..47S}. %Strong et al }
\caption{\gray longitude and latitude profiles, for model with pulsar source distribution and Xco varying with radius \cite{2004A&A...422L..47S}. %Strong et al Light blue line: bremsstrahlung, green line: inverse Compton scattering, red line: $\pi^0$-decay, black line: extragalactic background, dark blue line: total. }
\caption{The temperature dependences of the real ($\chi^{\prime}$) and imaginary ($\chi^{\prime \prime}$) ac susceptibilities of the ZFC dark states, i.e.~primordial states, are shown for all four batches of nanoparticles. All batches were measured with no applied static field and an alternating field of 4~G, except for Batch~D, which was measured in 1~G. The frequency dependence was studied at 1~Hz (\textcolor{red}{$\fullsquare$}), 10~Hz (\textcolor{green}{$\fullcircle$}), 100~Hz ({$\Diamond$}), and 1~kHz (\textcolor{blue}{$\fulltriangle$}) for all batches, except for Batch~D, which has an additional measurement at 333~Hz~($\ast$). Arrows are guides for the eyes and pass through the peaks.}
\caption{(Color online) Fractal scaling analysis for the two dimensional square lattice with the box-covering algorithm. Shown are the result of one Monte Carlo trial (\textcolor{red}{$\circ$}) and that obtained from 20 Monte Carlo trials (\textcolor{blue}{$\square$}). From the least-square-fit of the data (straight line), the fractal dimension is measured to be $\approx$2.0 as expected.}
\caption{(Color online)Fractal dimension of the Sierpinski gasket with the 12th generation measured by using the RS box-covering method. Solid line is guideline with slope $-\ln 3/\ln 2$. Shown are the result of one Monte Carlo trial (\textcolor{red}{$\circ$}) and that obtained from 20 Monte Carlo trials (\textcolor{blue}{$\square$}).}
\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}{$\circ$}) for the world-wide web. The results obtained from the two box-covering methods applied to the world-wide web are plotted here. The two methods yield the same fractal dimension $d_B\approx 4.1$. The method introduced by Song {\it et al.} is more optimal than ours in the viewpoint that $N_B(\ell_S) < N_B(2\ell_B+1)$. }
\caption{(Color online) Fractal scaling analysis for the world-wide web by the two box-covering algorithms, that of Song {\it et al.}\(\textcolor{blue}{$\square$}) and of ours (\textcolor{red}{$\circ$}). For comparison, the horizontal scale for that of Song {\em et al.}\is rescaled as$\ell_S/1.5 \to \ell_B$, by which we get the overlap of two curves obtained from the different algorithms. }
\caption{(Color online) Fractal scaling analysis for the world-wide web with the RS box-covering algorithm (\textcolor{red}{$\circ$}) and its variant that disallows disconnected boxes (\textcolor{blue}{$\square$})}
\caption{(Color online) Fractal scaling analysis for the square lattice (a) and the Sierpinski gasket (b) with the RS box-covering algorithm (\textcolor{red}{$\circ$}) and its variant that disallows disconnected boxes (\textcolor{blue}{$\square$}). Solid lines are guidelines with slopes of $-2$ in (a) and $-\ln 3/\ln 2$ in (b).}
\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) Fraction $F(f)$ of the vertices counted $f$ times in the cluster-growing algorithm for the Sierpinski gasket with the 12th generation composed of 265,721 vertices. $F(f)$ follows a Poisson-type distribution. 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) Fractal scaling analysis with the rules that allow (\textcolor{red}{$\circ$}) or disallow (\textcolor{blue}{$\square$}) the already-box-burned vertices to be chosen as the roots of new boxes for the world-wide web (a), and the fractal model (b).}
\caption{(color online) (a) Dumping the population initially in $\left|L\right>$ into highly excited, possibly unbound bands. (b) The resulting TOF image: the population initially in $\left|R\right>$ appears in the first BZ of the $\lambda$-lattice; the population initially in $\left|L\right>$ appears in higher BZs as a ring around the first BZ. (c) Calibration of the BS in Fig.~\ref{MZ}. After applying the BS at a given value of $\delta x/\lambda$, the population in the $\left|L\right>$ (red) or $\left|R\right>$ (blue) sites are measured by dumping the other site and measuring the fraction of atoms remaining in the first BZ. The value of $\delta x/\lambda$ where the two curves cross corresponds to a 50/50 BS. }% \label{dumpedBZ}% \end{figure} We measure the populations in the two output ports, $\left|L\right>$ and $\left|R\right>$, by selectively imparting energy to atoms in one of the two sites, which separates the populations in TOF (Fig.~\ref{dumpedBZ}). Transforming to the $\lambda$-lattice (in $300\,\mu$s) with its minimum centered over the $\left|R\right>$ site, imparts energy to $\left|L\right>$ atoms; $\left|R\right>$ atoms remain in the ground state of the $\lambda$-lattice (Fig.~\ref{dumpedBZ}a). The final states' Brillouin zones (BZs), which are filled, are observed (see for instance \cite{Kastberg1995,SebbyStrabley2006}) by turning off the lattice in 500\,$\mu$s and imaging after TOF (Fig.~\ref{dumpedBZ}b). The atoms in the center of the image correspond to the filled first BZ of the $\lambda$-lattice \cite{SebbyStrabley2006} and were originally in $\left|R\right>$. The atoms in the ring correspond to higher BZs and were originally in $\left|L\right>$. This method allows for \emph{direct measurement} of the atom population in $\left|L\right>$ and $\left|R\right>$. Using this method we can accurately calibrate the relative position of the two lattices that gives equal splitting of the NI-BS (shown in Fig.~\ref{dumpedBZ}c). This 50/50 BS was used for all experiments described here. To prepare number squeezed states with $N=1$ atom in $\left|g\right>$ of the $\lambda$-lattice sites, we used a two-stage ``slow" loading procedure. We started with $\approx 2\times 10^4$ atoms from a BEC with no discernable uncondensed fraction in a magnetic trap with $\omega_{\perp}/2\pi = 24$\,Hz and $\omega_{\parallel}/2\pi = 8$\,Hz and Thomas Fermi (TF) radii of $ \approx 13\,\mu$m and $ \approx 32\,\mu$m, respectively. The vertical lattice was slowly raised in 250\,ms, forming an array of 2D ``pancakes." After the first 100\,ms of this loading, we raised the $\lambda$-lattice in the remaining 150\,ms \footnote{We load to $\simeq 12\,E_R$ in 50\,ms then to $\simeq 100\,E_R$ in 100\,ms.}. %, a %timescale adiabatic with respect to lattice vibration and %interaction energies.}. % Ideally, at zero temperature, sufficiently slow loading produces a vertical array of 2D Mott insulating systems with one atom in $\left|g\right>$ of each $\lambda$-site. To prepare Poisson states with $\left<N\right> \approx 1$ in $\left|g\right>$, we used the same procedure, except we loaded both lattices quickly (500\,$\mu$s). This ``fast" load is adiabatic only compared to band excitation \cite{SebbyStrabley2006}. We destroyed the phase coherence between atoms in different $\lambda$-sites by holding atoms in the lattice for 5 ms. After preparing either of these initial states, we applied the NI-BS by transforming to the $\lambda/2$-lattice in 300 $\mu$s, a timescale fast enough to maintain coherence within a double well, yet slow enough to avoid vibrational excitation. After a variable hold time $t$, the atoms were released and imaged after 13 ms TOF. \begin{figure}[ptb] \includegraphics[scale=.50]{new2dslit_v7.eps}\caption{(color online) (a) A double-slit interference pattern (inset) and integrated density profile after 13\,ms TOF. (b) Example $y$-integrated diffraction pattern as a function of time for tilts $V=4.90(3)\,$kHz (left) and $V=-4.90(3)\,$kHz (right). (c) $C(t)$ for $N=1$ (black $\circ$), $N=2$ (green $\times$), and $\left<N\right>=1$ (blue $\blacksquare$). Solid lines are a fit to Eq.~\ref{ct}. Note: The plotted $C(t)$ for $N=2$ was scaled by 2. (d) $\phi(t)$ for the cases shown in (c). The $\phi(t)$ points which were derived from a signal with $2\times C(t) < 0.1$ are shaded in light green. $\phi(t=0)=0$ for all three cases, but each has been shifted for visual clarity. }% \label{dslit}% \end{figure} Fig.~\ref{dslit}a shows an example double-slit interference pattern. We extract the fringe visibility $C(t)$ and the spatial phase $\phi(t)$ by fitting the profile to: \begin{equation} F(x,t)=Ae^{-\frac{(x-x_{0})^2}{2\sigma^{2}}} \left[ 1+ C\left( t\right) \cos\left( \frac{x-x_{0}}{\Delta} + \phi(t)\right) \right],\nonumber \end{equation} where $\sigma$ is determined by the width of the ground state wavefunction in a $\lambda$-site, and $2 \pi \Delta$ is the fringe spacing. We use $C(t)$ to measure the number statistics of our input states. $\phi(t)$ evolves linearly due to an energy offset (tilt) $V$ between the sites within a double well % \footnote{ $V$ can result from misalignment of the $\lambda/2$-lattice~\cite{SebbyStrabley2006}, which we can in principle compensate with a small amount of the $\lambda$-lattice. The residual $\approx 1\,$kHz tilt in Fig.~\ref{dslit}d had no effect on the results presented here.} (see Fig.~\ref{dslit}b). % %Fig.~\ref{dslit}b shows example interference fringes %%as a function of $t$ %for two different tilts. % %an example positive and negative tilt of BLAH and BLAH. % %a positive and negative tilt. The NI-beam splitter binomially splits $N$ atoms in each $\lambda$-lattice site into superpositions of states $\left|n_L,n_R\right>$, with $n_L$ atoms in $\left|L\right>$ and $n_R=N-n_L$ atoms in $\left|R\right>$, resulting in the time dependent wavefunction \begin{equation}\label{wf} \left|\psi(t)\right>= \sum_{n_L=0}^N c(n_L, n_R)\left|n_L,n_R\right>e^{-i \omega(n_L,n_R) t} \end{equation} where $\left|c(n_L,n_R)\right|^2 = \left(\frac{1}{2}\right)^N \left(N!/n_L!n_R!\right)$. In the two-mode approximation interactions and tilt determine $\hbar \omega(n_L,n_R) = n_L V + (1/2) U n_L(n_L-1) + (1/2) U n_R(n_R-1)$, where $U$ is the on-site interaction energy per particle. For our system $U/h \approx 3$ kHz. In the absence of interactions and inhomogeneities, $\left|\psi\right>$ always factors into products of single particle states, $\left|\psi\right> = \left(\left|1,0\right>+e^{i\omega_{\rm{tilt}}t}\left|0,1\right>\right)^N$, where $\omega_{\rm{tilt}} = V/\hbar$. All the terms in Eq.~\ref{wf} constructively interfere and $C(t)$ is maximized and constant. When interactions are included, $\left|\psi\right>$ is not always factorable, and $C(t)$ (for atoms in the ground band) is modulated by $\left|\cos^{N-1}( U t/\hbar)\right|$ \cite{Johnson2006,WALLS1997}. Summed over a distribution of site occupation numbers, the total $C(t)$ is given by \begin{equation}\label{ct} C(t) = \frac{e^{-\Gamma t}}{\sum_{N=1}^{N_{\rm{max}}} f_N N}\left|\sum_{N=1}^{N_{\rm{max}}} f_N N \cos^{N-1}(U t/\hbar)\right|\end{equation} where $f_N$ is the fraction of $\lambda$-sites with $N$ atoms (for our case we limited $N_{\rm{max}}=4$), $\sum f_N = 1$, and $\Gamma$ is %----Referee 1.3 change: an empirical dephasing rate that %depends on the lattice alignment and that we attribute to lattice inhomogeneities. %---- Fitting $C(t)$ with Eq. \ref{ct}, we can extract the relative fractions $f_N$. For the slow loading case with $N_{\rm{BEC}}\leq 2\times 10^4$ (average filling $\leq 1$ per site) we expect to have a Mott state; we measure $C(t)$ to be approximately constant (see Fig. \ref{dslit}c), indicating that occupied $\lambda$-lattice sites have $N=1$. ($C(t)$ shows a slow decay, which we attribute to tilt inhomogeneities.) Extracting $f_N$ from $C(t)$, we determine $f_1=0.94(6)$. (As shown in Fig.\,\ref{mott}a, $f_1$ was optimized at a load time of 150\,ms.) The uncertainty in$f_1$ is dominated by the shot-to-shot scatter in the double-slit visibility $C$, which is significantly larger than for fast loading. We do not understand this increased noise, but note that the shot-to-shot scatter in $\phi$ (Fig.\,\ref{dslit}d) is not increased. This suggests an increased sensitivity to an uncontrolled initial condition, e.g., number or temperature. The near-zero value of $f_2$ indicates a strongly number-squeezed initial state, as expected in a Mott insulator (compared to the Poisson value of $0.3$ for average filling of unity, ignoring $f_0$). The value of $f_1$ does not, however, necessarily represent the fidelity of the Mott state since this measurement is insensitive to sites with $N=0$. Assuming $f_1=0.94$ and $f_2=0.06$ and a homogeneous system, we can roughly estimate the temperature $T$ in the $\lambda$-lattice from $f_2=e^{-U/k_B T}$. This gives $T \leq 0.36 U/k_B\simeq50$\,nK. For the fast loading case where we expect a Poisson distribution of site occupation numbers, there can be$N>1$ atoms at any $\lambda$-site. In this case we find that $C(t)$ shows distinct collapses and revivals, % as a function of time, caused by interactions (see Fig.~\ref{dslit}c). Only at integer multiples $l$ of $T_{\mathrm{rev}}=h/U$ does the wavefunction within a site factor: $\psi = \left( \left|1,0\right> +e^{il\omega_{\rm{tilt}}T_{\rm{rev}}} \left|0,1\right>\right)^N$. We observe revivals in $C(t)$ with a period of $\approx 350\,\mu$s, in good agreement with our calculated value of $U/h = 2.88$\,kHz. While similar to the interaction-induced collapse and revival seen in\cite{Greiner2002b}, the interference here is between the sites of a double well and is created by a topology change in the lattice, rather than between sites extending over the entire lattice and created by changing the lattice depth. This allows us to see interference even in the absence of global coherence, as evidenced by the interference pattern generated from the highly squeezed $N=1$ slow load state. \begin{figure}[ptb] \includegraphics[scale=.50]{new2f1vsloadv5_bw.eps}\caption{We use Eq.~\ref{ct} to fit $C(t)$ for the fast and slow load data. (a) Extracted $f_1$ for varied slow load times. The uncertainties are from the fit to Eq. \ref{ct}. (b) Using Eq.~\ref{ct} to fit the fast load $C(t)$ we extract values for $f_1,f_2,f_3$ and $f_4$. From those coefficients we determine $\left<N\right>$, the peak number density of atoms in $\lambda$-lattice sites, as a function of $N_{\mathrm{BEC}}$. The line and the grey shaded area are the peak number density scalings predicted by the TF-approximation (with no adjustable parameters) and our estimated uncertainty in the calibration of $N_{\mathrm{BEC}}$ (factor of $\pm 1.5$). The uncertainties on the individual points are from a combination of the uncertainties in $f_i$ and from the fit determining $\left<N\right>$.}% \label{mott}% \end{figure} Using Eq.~\ref{ct} to fit the fast load $C(t)$, we extract the number distribution of atoms in $\lambda$-lattice sites. By fitting that distribution to an average over Poisson distributions whose local mean occupation numbers are proportional to the initial TF-density profile, we extract $\left<N\right>$, the peak number density of atoms in $\lambda$-lattice sites. % This weighted Poissonian distribution fits the data within experimental uncertainty. % Fig. \ref{mott}b shows $\left<N\right>$ as a function of $N_{\mathrm{BEC}}$ % %. Also plotted is % and the power law scaling ($N_{\rm{BEC}}^{2/5}$) for $\left<N\right>$ expected from the TF-approximation. % %(with no adjustable parameters) % For large values of $N_{\mathrm{BEC}}$ the data deviates from the trend set by the lower values of $N_{\mathrm{BEC}}$. While not completely understood, this result may be attributed to three-body loss mechanisms and a larger thermal fraction for larger values of $N_{\mathrm{BEC}}$. % if we had room I'd like to refer to Jaksch, Cirac, and Zoller AND % a reference to the power law scaling of the TF dist. We now consider input states with $N=2$ atoms in $\left|g\right>$ of the $\lambda$-lattice sites. If we simply loaded the $\lambda$-lattice from our harmonic trap with increased initial peak density, the resulting ``shell structure" should produce a significant number of sites with $N=1$ surrounding the central core of $N=2$ \cite{Folling2006, Campbell2006}, limiting the maximum value of $f_2$. In principle, the flexibility of the double well lattice can be used to construct a purer $N=2$ state in the $\lambda$-lattice by combining two $\lambda/2$-lattice sites: we first load the vertical lattice as in the slow load case above, but with an initial number ($N_{\mathrm{BEC}} \approx 7\times 10^4$) large enough to produce $\simeq 1$ atom per site in the $\lambda/2$-lattice. We then raise the $\lambda/2$-lattice in 100\,ms (ideally, slow enough to make a Mott insulating state with one atom in the ground state of each$\lambda/2$-site). Finally we adiabatically (in $30$\,ms) combine pairs of$\lambda/2$-sites into $\lambda$-sites, thus performing an I-BS. The input states are twin Fock states of one atom each in $\left|L\right>$ and $\left|R\right>$, %---Referee 2.4 the two-atom interacting ground state of the $\lambda/2$-lattice. %---Referee 2.4 The output modes are $\left|g\right>$ and $\left|e\right>$ states of $\lambda$-sites. Interactions and adiabaticity % ensure that both atoms go into $\left|g\right>$, the two-atom ground state of the $\lambda$-lattice. %--- This ``constructed pair" state is then input to our interferometer (split with an NI-BS). As shown in Fig.~\ref{dslit}c,d we find that after using the constructed pair technique to load the $\lambda$-lattice, $C(t)$ and $\phi(t)$ show unique features which can only be attributed to $\lambda$-sites with $N=2$. As in the fast load case, $C(t)$ collapses and revives, but there are revivals at half the period compared to the fast load case ($T_{\mathrm{rev}}/2$ rather than $T_{\mathrm{rev}}$). Like the slow and fast load cases, $\phi(t)$ evolves linearly as $\omega_{\rm{tilt}}$ (here $\approx 0$), but here $\phi(t)$ makes clear jumps between revivals. These jumps are $\simeq\pi$, particularly if only the $\phi(t)$ points with $2 \times C(t) > 0.1$ are considered. At $lT_{\mathrm{rev}}$ the wavefunction can be factored as $\left|\psi\right>=\left(\left|1,0\right>+ e^{i\omega_{\rm{tilt}}T_{\mathrm{rev}}} \left|0,1\right>\right)^2$, so $C(t)$ revives. Unlike the fast load case, at $lT_{\mathrm{rev}}/2$, $C(t)$ also revives but with the wavefunction factored into oppositely-phased single particle states, $\left|\psi\right>=\left(\left|1,0\right>- e^{i\omega_{\rm{tilt}}T_{\mathrm{rev}}/2}\left|0,1\right>\right)^2$, giving a $\pi$ phase-shifted interference pattern~\cite{Johnson2006,WALLS1997}. %We were never %able to see the $T_{\mathrm{rev}}/2$ revival by merely loading into %the $\lambda$-lattice with higher density; only with the constructed %pairs technique was this feature observed. Our maximum measured $C$ in Fig.~\ref{dslit}c for the constructed pairs technique is a factor of $\approx$ 2 less than the fast and slow load cases, and our measured $C(T_{\mathrm{rev}}/2)$ is lower than $C(T_{\mathrm{rev}})$. %---Referee 1.4 These differences % can be largely explained % % may be explained in part % by the fact that after the I-BS, a large fraction of atoms ($\simeq$ 30$\%$) are in $\left|e\right>$ (revealed by imaging the $\lambda$-lattice BZ), % and make an interference pattern $\pi$-out-of-phase with that made by atoms in $\left|g\right>$ % \footnote{ % %This requires a modification to Eq.~\ref{ct}, %which does not account for $\left|e\right>$ population % Eq.~\ref{ct} does not account for $\left|e\right>$ population. % }
\caption{\label{Fig_Saturation}\textbf{a} The Rydberg atom number plotted versus excitation time for a high laser intensity ($\Omega_0=$ 210 kHz) and three values of the density of ground state atoms $n_{\text{g,0}}= $ (\unit[3.2$\times 10^{13}$] (\textcolor{green}{$\diamond$}), \unit[6.6$\times 10^{12}$] (\textcolor{red}{$\triangledown$}), \unit[2.8$\times 10^{12}$] (\textcolor{blue}{$\square$})) {cm$^{-3}$}. \textbf{b} The Rydberg atom number plotted versus excitation time for a low atom density ($n_{\text{g,0}}=2.8 \times 10^{12}\text{cm}^{-3}$) and three values of laser intensity $\Omega_0= $ (\unit[210] (\textcolor{blue}{$\square$}), \unit[93] (\textcolor{red}{$\vartriangle$}), \unit[42] (\textcolor{green}{$\circ$}) ) kHz. The solid curves are fits to the data with a simple exponential saturation curve. The inset in \textbf{a} shows a magnification of the data in contrast with the calculated Rabi oscillation (dashed) assuming negligible interactions. Fig. \textbf{c} shows a schematic of the excitation dynamics in an inhomogeneous sample. Many oscillating `superatoms' (shown with exaggerated amplitudes in \textcolor{red}{red}) add up to an integrated staturation curve (\textcolor{blue}{blue}). This curve falls behind the noninteracting case (shown for short times dashed in \textcolor{blue}{blue}) on a time scale of less than \unit[50]{ns}. }
\caption{\label{Fig_slope}\textbf{a} Dependence of the initial increase $R$ of the Rydberg atom number on the density of ground state atoms $n_{\text{g,0}}$ for high (\textcolor{red}{$\square$}) and low (\textcolor{blue}{$\circ$}) Rabi frequency $\Omega_0= $(210, 42)kHz. \textbf{b} Dependence of $R$ on the excitation rate $\Omega_0$ for high (\textcolor{black}{$\vartriangle$}) and low (\textcolor{green}{$\diamond$}) atom density $n_{\text{g,0}}= $(7.2$\times10^{13}$, 2.8$\times10^{12}$) cm$^{-3}$. %The red dataset (\textcolor{red}{$\square$}) contains the measurements shown in Fig. \ref{Fig_Saturation}a, % the green dataset (\textcolor{green}{$\diamond$}) contains the measurements shown in Fig. \ref{Fig_Saturation}b. The lines are the result of a power-law fit to the whole dataset in \textbf{a} and \textbf{b} of the form $R\propto n_{\text{g,0}}^{a} \Omega_0^{b}$ which gives an exponent for the $n_{\text{g,0}}$-dependence of $a=0.49\pm$0.06 which is in excellent agreement with the expected $\sqrt{n_{\text{g,0}}}$-scaling for collective excitation. The fitted exponent for the $\Omega_0$-scaling is $b=1.1\pm$0.1, which is in good agreement with a linear scaling with $\Omega_0$ for coherent excitation.}
\caption{\label{Fig_satvalue}\textbf{a} Dependence of the saturation number of Rydberg atoms $N_{\text{sat}}$ on the density of ground state atoms $n_{\text{g,0}}$ for high (\textcolor{red}{$\square$}) and low (\textcolor{blue}{$\circ$}) Rabi frequency $\Omega_0=$(210, 42) kHz. \textbf{b} Dependence of $N_{\text{sat}}$ on the Rabi frequency $\Omega_0$ for high (\textcolor{black}{$\vartriangle$}) and low (\textcolor{green}{$\diamond$}) atom density $n_{\text{g,0}} = $ (7.2$\times10^{13}$, 2.8$\times10^{12}$) cm$^{-3}$ . %The red dataset (\textcolor{red}{$\square$}) contains the measurements shown in Fig. \ref{Fig_Saturation}a, % the green dataset (\textcolor{green}{$\diamond$}) contains the measurements shown in Fig. \ref{Fig_Saturation}b. The lines are the result of a power-law fit to the whole dataset in \textbf{a} and \textbf{b} of the form $N_{\text{sat}}\propto n_{\text{g,0}}^{c} \Omega_0^{d}$ which gives an exponent for the $n_{\text{g,0}}$-dependence of $c=0.07\pm$0.02, which is in agreement with the expected independence from $n_{\text{g,0}}$ for strong blockade. The fitted exponent for the $\Omega_0$-scaling is $d=0.38\pm$0.04, which is in excellent agreement with the expected $\Omega_0^{2/5}$-scaling for a collective van der Waals blockade.}
\caption{Canonical phase diagram for $\Gamma XW$ spin spirals in the fcc structure. }{\includegraphics[scale=0.8,angle=00]{pdGXW-1.eps}}
\caption{Canonical phase diagram for $\Gamma HP$ spin spirals in the bcc structure. }{\includegraphics[scale=0.8,angle=00]{pdGHP-1.eps}}
\caption{Canonical spin spiral spectrum's for the symmetry path $\Gamma$XWL$\Gamma$ in the fcc Brillouin zone. Shown are spectrums for (a) $n=9.3$, $m=0.5$, (b) $n=8.8$, $m=0.4$, (c) $n=7.5$, $m=1.2$, and (d) $n=0.5$, $m=3.3$. Shown in the lowest panel is (e) the region around the X point for $n=6.5$, $m=0.8$. The top scale refers to graphs (a)-(d), and the bottom to graph (e). \vspace{0.7cm} }{\includegraphics[scale=0.3,angle=00]{fcc-spec.eps}}
\caption{Canonical Density of States for the (a) ferromagnetic, (b) $\bq = [0.0,0.0,0.5]$ spin spiral, and (c) antiferromagnetic X-point structure. \vspace{1.0cm} }{\includegraphics[scale=0.3,angle=00]{dos.eps}}
\caption{Energy difference of FM and $\bq = [0.0,0.0,0.1]$ spin spiral structures. Full line is total energy different whilst caption indicates components. \vspace{1cm} }{\includegraphics[scale=0.3,angle=00]{fm-GX0.1.eps}}
\caption{ Dalitz plot of Christ's variables $T_c^\ast$ and $E_{\gamma}^\ast$ and contour plot of the \red{ Inner Bremsstrahlung} amplitude: the right corner (in red) has a greater density. }
\caption{Values of $A^+$ and $\eta_V$ on the \red{ central solid line} generate the \red{E787} experimental value of the branching ratio in TAB. \ref{tab:kppg} while the other lines are the borders of strips corresponding to{\color{arancia}{ $1\sigma$}}, {\color{verde}{$2\sigma$}} and {\color{bluazz}{$3\sigma$}} deviation from the E787 central value in TAB. \ref{tab:kppg}. As a result the {\color{blue}{lower solid line}} corresponds to the central {\color{blue} NA48} value in TAB. \ref{tab:kppg}. For $\eta_V=0$ and $A^+=2$, the amplitude is dominated by the WZW pole. }
\caption{Normalized $T_c^*$-spectra ($T_{c}^{\ast }\in \left[0, 80\right]$ MeV) for the DE magnetic amplitude with the E787 branching ratio \cite{BNL00} in TAB. \ref{tab:kppg} (upper plot) and the NA48/2 branching ratio in TAB. \ref{tab:kppg} (lower plot). The \red{solid} curves correspond to a \red{constant} amplitude, the {\color{blue} dotted} curves to a magnetic form factor with {\color{blue} $\eta_V=1.5$} and corresponding value of $A^+$ on the central (lower) full line of FIG. 2 for the upper (lower) plot . }}\label{fig:TcNA48}\vskip.25cm \end{figure} In FIG. 3 we show the normalized $T_c^*$-spectra for the DE magnetic amplitude with the E787 branching ratio \cite{BNL00} in TAB. \ref{tab:kppg} (upper plot) and the NA48/2 branching ratio in TAB. \ref{tab:kppg} (lower plot). The solid curves correspond to a constant amplitude, the dotted curves to a magnetic form factor with $\eta_V=1.5$ and corresponding value of $A^+$ on the central (lower) full line of FIG. 2 for the upper (lower) plot. Then we plot the $W$-spectra with $T_{c}^{\ast }\in \left[55, 90\right]$ MeV in FIG. 4 and $T_{c}^{\ast }\in \left[0, 80\right]$ MeV in FIG. 5. In each case we consider form factors with $\eta_V=0.5$ (dashed curves) and $\eta_V=1.5$ (dotted curves); the corresponding values of $A^+$ are determined from FIG.2. \begin{figure}[hb] \vskip.25cm\centering \epsfysize=6cm\epsfxsize=7cm\epsfbox{plotWAdler5590.eps} \epsfysize=6cm\epsfxsize=7cm\epsfbox{plotWNA485590.eps} \smallskip \parbox{7cm}{\caption{Normalized $W$-spectra ($T_{c}^{\ast }\in \left[55, 90\right]$ MeV) for the DE magnetic amplitude. The upper and lower figures correspond to the E787 branching ratio \cite{BNL00} in TAB. \ref{tab:kppg} (the central solid line of FIG. 2) and to the NA48/2 branching ratio in TAB. \ref{tab:kppg} (the lower solid line of FIG. 2) respectively. The solid curves corresponds to a \red{constant} amplitude, while the {\color{verdel} dashed} and {\color{blue} dotted} curves correspond to form factors with {\color{verdel} $\eta_V=0.5$} and {\color{blue} $\eta_V=1.5$} respectively. }
\caption{\label{fig:experimental-setup} \redtext{Schematic view of the experimental setup of CAST. The TPC detector which observes the sun during sunset is shown on the left side. On the right side the X-ray telescope system is shown. Please note that the Micromegas detector and the tracking system are not shown in this picture. The whole system is operated at a pressure of $\approx 10^{-6}\,\text{mbar}$}}
\caption{\label{fig:background-ccd} Left: Background spectrum observed during non-tracking times, i.e., while the CAST magnet was not pointing to the sun but under the same operating conditions as during observations of the sun \citeaffixed{andriamonje:07a}{see}. The fluorescent emission lines apparent in the spectrum are labeled. The energy range which is sensitive for axion detection is marked as \emph{Region of Interest}. Right: Background spectra observed with the pn-CCD under different shielding conditions. From top to bottom: Background observed with the internal copper \redtext{shield (black) and with the internal lead and copper shield (blue). The lowest spectrum (red) represents the observed background with the final shield configuration, which consists (from the outside to the inside) of an external lead shield followed by the evacuated detector vessel, and the internal internal lead and copper shield.}}
\caption{\label{fig:background-position} Left: Count rate observed with the X-ray telescope depending on the pointing direction of the magnet. Each cell has a dimension of $10\,\degree\times2\,\degree$. \redtext{The attitude of the X-ray telescope is given in topocentric horizontal coordinates.} % The numbers given in each cell indicates the integral observed count % rate over the full CCD area in the $1$--$7\,\text{\kev}$ energy range % and its uncertainty on the $00\,\%$ confidence level. Right: Integration time for each cell.}
\caption{(color online) Real ({\color{red} solid}) and imaginary ({\color{blue} dashed}) parts of the electric ($\alpha_{EE}$) and magnetic ($\alpha_{HH}$) polarizabilities as well as the chirality parameters ($\alpha_{HE}$, $\alpha_{EH}$) in arbitrary but the same units. % , as a function of % probe field detuning $\Delta=-\Delta_E=-\Delta_B$ relative to the % radiative decay rate $\gamma_3$ from level $|3\rangle$. Here, % $\Delta_E=\omega_{34}-\omega_{\rm probe}$ and % $\Delta_B=\omega_{21}-\omega_{\rm probe}$, where $\omega_{\rm probe}$ % is the probe-field frequency and $\omega_{\mu\nu}$ is the transition % frequency between levels $|\mu\rangle$ and $|\nu\rangle$. All % coupling fields are resonant. }
\caption{(color online) Real ({\color{red} solid}) and imaginary ({\color{blue} dashed}) parts of the refractive index % , including local field effects, % as a function of the probe field detuning $\Delta=-\Delta_E=-\Delta_B$ % relative to the radiative decay rate $\gamma_2$ from level % $|2\rangle$, for a density of (a) $N=5\cdot 10^{16}\text{cm}^{-3}$ and (b) $N=5\cdot 10^{17}\text{cm}^{-3}$. % Other system parameters are given in the text % and all coupling fields are resonant. }
\caption{(color online) Real ({\color{red} solid}) and imaginary ({\color{blue} dashed}, $\times 100$) parts of the refractive index, as well as the real part of the permeability ({\color{green} dotted}), as a function of the logarithm of the density $N$, at a frequency slightly below resonance ($\Delta=-25\gamma_2$). % and including local field effects. Inset: $\log\left|{\Re(n)}/{\Im(n)}\right|$ as a function of $\log[N]$. % Other system parameters are the same as for % Fig.~\ref{fig2} and all coupling fields are resonant. } \label{fig4} \end{center} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% It has been pointed out by Smith {\it et al.} and Merlin \cite{Schmith03} that the realization of sub-diffraction-limit imaging with a lens of thickness $d$ and resolution $\Delta x$ requires an extreme fine tuning of the index of refraction to the value $n=-1$ with accuracy $\Delta n = 1-\exp\left\{-\frac{\Delta x}{2\pi d}\right\}$. The quantum interference scheme presented here provides a handle for such fine tuning. For example, as shown in Fig.~\ref{fig-tuning}, the real part of the refractive index can be fine tuned by relatively coarse adjustments of the strength of the coupling field $\Omega_c$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[tb] \begin{center} \includegraphics[width=6cm]{Omegadependence.paper.PRL.eps} \caption{(color online) Real ({\color{red} solid}) and imaginary ({\color{blue} dashed}) parts of the refractive index as a function of the coupling field Rabi-frequency $\Omega_c$ % relative to the radiative decay rate $\gamma_3$, for $N=3.5\cdot10^{17}$ cm$^{-3}$ and $\Delta=-25\gamma_2$. % $\Delta=-\Delta_B=-\Delta_E=-25\gamma_2$. % Other system parameters are the same as for % Figs.~\ref{fig2} and \ref{fig4} and all coupling fields are resonant. } \label{fig-tuning} \end{center} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% As with EIT and its applications, electromagnetically-induced chiral negative refraction should be realizable in a wide range of atomic, molecular, and condensed matter systems. Already, a basic form of electromagnetically-induced chirality has been realized in Rb vapor \cite{Scully-chirality}. In preparation for further experimental investigations we are currently performing a detailed assessment of systems such as metastable neon, rare-earths (dysprosium and others), as well as donor-bound electrons and bound excitons in semiconductors, all of which can have level structures and interactions analogous to those of the modified level scheme of Fig.~\ref{fig1}(b). %%%%%%%%%%%%%%%% % Results from these systems will be reported in future % publications. %%%%%%%%%%%%%%%% In conclusion, we showed that quantum interference effects can lead to a large induced chirality under realistic conditions, and thus enable tunable negative refraction with minimal absorption. Such electromagnetically-induced chiral negative refraction should be applicable to a wide array of systems in the optical regime, including atoms, molecules, quantum dots, and excitons. \begin{acknowledgments} M.F. and J.K. thank the Institute for Atomic, Molecular and Optical Physics at the Harvard-Smithsonian Center for Astrophysics and the Harvard Physics Department for their hospitality and support. R.W. thanks D. Phillips for useful discussions. J.K. acknowledges financial support by the Deutsche Forschungsgemeinschaft through the GRK 792 ``Nichtlineare Optik und Ultrakurzzeitphysik''. S.Y. thanks the Research Corporation for support. \end{acknowledgments} \def{\}{etal}{\textit{et al.}} \begin{thebibliography}{99} \bibitem{Veselago68} V. G. Veselago, Sov. Phys. Usp. {\bf 10}, 509 (1968). \bibitem{Pendry00} J. B. Pendry, Phys. Rev. Lett. {\bf 85}, 3966 (2000). \bibitem{Pendry99} J. B. Pendry \etal, IEEE Trans. Micro. Theory Tech. {\bf 47}, 2075 (1999). \bibitem{Smith00} D. R. Smith \etal, Phys. Rev. Lett. {\bf 84}, 4184 (2000); R. Shelby, D. R. Smith, and S. Schultz, Science {\bf 292}, 77 (2001). \bibitem{Yen04} T. J. Yen \etal, Science {\bf 303}, 1494 (2004). \bibitem{Linden04} S. Linden \etal, Science {\bf 306}, 1351 (2004); C. Enkrich \etal, Phys. Rev. Lett. {\bf 95}, 203901 (2005). \bibitem{Parimi04} P. V. Parimi \etal, Phys. Rev. Lett. {\bf 92}, 127401 (2004). \bibitem{Berrier04} A. Berrier \etal, Phys. Rev. Lett. {\bf 93}, 073902 (2004). \bibitem{Lu05} Z. Lu \etal, Phys. Rev. Lett. {\bf 95}, 153901 (2005). \bibitem{Fleischhauer05} M. Fleischhauer, A. Imamoglu, and J. P. Marangos, Rev. Mod. Phys. {\bf 77}, 633 (2005). \bibitem{Scully-chirality} V. A. Sautenkov \etal, Phys. Rev. Lett. {\bf 94}, 233601 (2005). \bibitem{Schmith03} D. R. Smith \etal, Appl. Phys. Lett. {\bf 82} 1506 (2003); R. Merlin, Appl. Phys. Lett. {\bf 84}, 1290 (2004). \bibitem{Pendry04} J. B. Pendry, Science {\bf 306}, 1353 (2004). \bibitem{Harris-NLO-EIT} S. E. Harris, J. E. Field, and A. Imamoglu, Phys. Rev. Lett. {\bf 64}, 1107 (1990). \bibitem{Stoicheff-Hakuta} K. Hakuta, L. Marmet, and B. P. Stoicheff, Phys. Rev. Lett. {\bf 66}, 596 (1991). \bibitem{Oktel04} M. \"O. Oktel, and\"O. E. M\"ustecaplioglu, Phys. Rev. A{\bf 70}, 053806 (2004). \bibitem{Thommen-PRL-2006} Q. Thommen, and P. Mandel, Phys. Rev. Lett. {\bf 96}, 053601 (2006). \bibitem{footnote} In \cite{Thommen-PRL-2006} a negative value for Im$[n]$ is given, corresponding to an amplifying medium; however this result is due to a calculational error, see J. K\"astel,\& M. Fleischhauer, Phys. Rev. Lett.{\bf 98}, 069301 (2007). \bibitem{Cook} D. M. Cook, {\it The Theory of the Electromagnetic Field}, Prentice-Hall (New Jersey). \bibitem{Kaestel07} For a rigorous derivation of generalized Clausius-Mossotti relations in magneto-dielectric media see: J. K\"astel, G. Juzeliunas, and M. Fleischhauer (in preparation)\end{thebibliography} \end{document}}
\caption{(color online) Real ({\color{red} solid}) and imaginary ({\color{blue} dashed}) parts of the refractive index as a function of the coupling field Rabi-frequency $\Omega_c$ % relative to the radiative decay rate $\gamma_3$, for $N=3.5\cdot10^{17}$ cm$^{-3}$ and $\Delta=-25\gamma_2$. % $\Delta=-\Delta_B=-\Delta_E=-25\gamma_2$. % Other system parameters are the same as for % Figs.~\ref{fig2} and \ref{fig4} and all coupling fields are resonant. }
\caption{The $<A_1>$ values in the 1.5 to 2.5 K$^\prime$-band scale length of Ursa Major galaxies. The Eridanus Group values are taken from Paper I. The values denoted by \otriangle, correspond to A$_1$ values of the 3 galaxies estimated from R-Band analysis (see Discussion ).The A$_1$ values are higher for the early-type galaxies in the Eridanus while the distribution is flatter for the Ursa Major group.}
\caption{Lattice parameters of PZT20/80 bulk (c$_b$ \textcolor{red}{$\triangle$}, a$_b$ \textcolor{red}{$\triangledown$}) and thin films (c$^{\perp}_f$ \textcolor{blue}{$\blacktriangle$}, a$^{\perp}_f$ \textcolor{blue}{$\blacktriangledown$}) deposited on MgO ($\vardiamondsuit$). Inset: High-temperature evolution of the out-of-plane lattice parameter. Full line arises from a calculation based on an elastic behavior of the film (see text).}
\caption{Some of the instanton numbers $n_{(n_1,n_2,n_3,m_1,m_2)}$ computed by mirror symmetry. The entries marked in \textcolor{red}{\textbf{bold}} depend non-trivially on the torsion part of their respective homology class.}
\caption{\label{fig:1}Energy-level scheme of three qubits, with $ \Delta_{\pm} = -\Delta_{13} \pm \frac{1}{2}\left(\Delta_{13} + \Omega \right)$, for $\Omega \equiv \sqrt{8 \Delta_{12}^2 + \Delta_{13}^2}$. In the Dicke limit~\cite{Dicke}, $\ket{\sf b} \equiv \tfrac{1}{\sqrt{6}} (-2\ket{001} + \ket{010} + \ket{100})$ and $\ket{\sf c} \equiv \tfrac{1}{\sqrt{2}} (\ket{010} - \ket{100}) $. The DFS is labeled $\{ \ket{0}_L,\ket{1}_L \}$~\cite{Kem01}.} \end{figure} The laser has a bichromatic electric field $\boldsymbol{E}(\boldsymbol{r}) = \boldsymbol{E}_\mu(\boldsymbol{r}) + \boldsymbol{E}_\nu(\boldsymbol{r})$, with $\boldsymbol{E}_\mu(\boldsymbol{r})$ and $\boldsymbol{E}_\nu(\boldsymbol{r})$ the electric field amplitudes. $\boldsymbol{E}(\boldsymbol{r})$ interacts with the three qubits via a dipole coupling and so we introduce the Rabi frequencies $\mathcal{E}_{\mu,i} = \vec{d} \cdot \boldsymbol{E}_\mu \text{e}^{-\text{i} \boldsymbol{k}_\mu \cdot \boldsymbol{r}_{i}}$ and $ \mathcal{E}_{\nu,i} = \vec{d} \cdot \boldsymbol{E}_\nu \text{e}^{-\text{i} \boldsymbol{k}_\nu \cdot \boldsymbol{r}_{i}} $ for wave vectors $\boldsymbol{k}_\mu$ and $\boldsymbol{k}_\nu$, and where qubit $i$ is situated at $\boldsymbol{r}_i$. Within the rotating-wave approximation, the interaction Hamiltonian is written \begin{align} \widehat{H}_{\text I} = \sum^{3}_{i=1} \mathcal{E}_{i} \widehat{\sigma}_{i-} + \text{H.c.}, \end{align} for which $\mathcal{E}_{i}$ are copropagating and described by a time-dependent bichromatic external field, \begin{align} \mathcal{E}_{i} = \mathcal{E}_{\mu,i} \text{e}^{-\text{i}\omega_\mu t } + \mathcal{E}_{\nu,i} \text{e}^{-\text{i}\omega_\nu t }. \end{align} Due to the different positions of the qubits, the field $\mathcal{E}_{i}$ differs for distinct qubits. The total effective Hamiltonian for the no-jump evolution is $\widehat{H}_{\text{eff}} = \widehat{H}_{\text S} + \widehat{H}_{\text I}$. The jump operators are identified by diagonalizing the relaxation matrix $(\gamma_{ij})$~\cite{Clem03}, \begin{align} (\gamma_{ij}) = \boldsymbol{B}^{T} \boldsymbol{\Lambda} \boldsymbol{B}, \end{align} where $\boldsymbol{\Lambda} \equiv \textrm{diag}\left(\lambda_1,\lambda_2,\lambda_3\right)$ is a diagonal matrix of the eigenvalues of $(\gamma_{ij})$ and the columns of $\boldsymbol{B}^{T} = (b_{ij})^{T}$, \begin{align} \boldsymbol{b}_{l} = \begin{pmatrix} b_{l1} \\ b_{l2} \\ b_{l3} \end{pmatrix} \end{align} are the corresponding normalized eigenvectors. We define $\boldsymbol{\widehat{\Sigma}}^{\dagger} \equiv \left( \widehat{\sigma}_{1+},\widehat{\sigma}_{2+} , \widehat{\sigma}_{3+} \right)$, so the jump operators are written \begin{align} \widehat{J}_{l} = \sqrt{\lambda_{l}}\boldsymbol{b}^{T}_{l}\widehat{\boldsymbol{\Sigma}} \qquad \textrm{and} \qquad \widehat{J}^{\dagger}_{l} = \sqrt{\lambda_{l}} \boldsymbol{\widehat{\Sigma}}^{\dagger} \boldsymbol{b}_{l}. \end{align} These are quoted explicitly for three qubits in Ref.~\cite{Clem03}. In this unraveling, the master equation is \begin{align} \label{eq:me} \dot{\hat{\rho}} = -\frac{\text{i}}{\hbar}(\widehat{H}_{\text{eff}}\hat{\rho} - \hat{\rho} \widehat{H}^{\dagger}_{\text{eff}}) + \sum_{i=1}^{3} \widehat{J}_{i}\hat{\rho}\widehat{J}_{i}^{\dagger}, \end{align} which corresponds to the standard Lehmberg master equation for three qubits~\cite{Bela69Lehm70iLehm70iiArg70} and is simply the Lindblad master equation~\cite{Lind}. This unraveling is useful for analyzing preparation, manipulation, and read out of DFS-encoded quantum information in dipole-coupled qubits. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Preparation} \label{sec:prep} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% We describe how to prepare the maximally entangled DFS state $\ket{\sf b}$ deterministically. This state is the lower state of $\ket{0}_L$, and is the longest-lived excited state in the eight-level system. For the purposes of preparation, we require only a single laser field, \begin{align} \label{eq:slf} \widehat{H}_{\text I} = \sum^{3}_{i=1} \mathcal{E}_{\mu,i} \text{e}^{-\text{i}\omega_\mu t } \widehat{\sigma}_{i-} + \text{H.c.}, \end{align} with frequency $\omega_\mu$, and $\boldsymbol{k}_\mu$ orthogonal to the line joining the qubits. \begin{figure}[tbp] \begin{center} \subfigure[]{\label{fig:2a} \includegraphics[width=3.5cm,height=2.5cm]{fig2a.eps}} \subfigure[]{\label{fig:2b} \includegraphics[width=3.5cm,height=3.0cm]{fig2b.eps}} \end{center} \caption{(a) \label{fig:2} Population of the levels {\sf a,\ldots,h} for $\xi_{12} = 0.5$, $\mathcal{E}_\mu = \gamma$, $\omega_\mu = \tfrac{1}{2}(\Delta_{13} - \Omega )$, and $\alpha = 0$. For these parameters, $F=0.988$ and the time taken for the transfer {\sf a-b}: $t_{\pi} = 0.987 \gamma^{-1}$. (b) Number of population inversions possible during time period $\gamma_{\sf b}^{-1}$ versus qubit separation with $F = 0.98$ and $\alpha =0$. For the separation proposed in Ref.~\cite{Me05a}, $\gamma_{\sf b} t_\pi \approx 5 \times 10^{-7}$. } \end{figure} We assume that the initial state is the ground state and that the qubits are positioned according to $\boldsymbol{r}_{1}= - \boldsymbol{r}$, $\boldsymbol{r}_{2}= 0$, and $\boldsymbol{r}_{3}= \boldsymbol{r}$. Under these conditions, in the interaction picture with respect to the Hermitian part of $\widehat{H}_{\text S }$, the effective coupling between states $\ket{\sf a}$ and $\ket{\sf b}$ in the collective basis is \begin{align} \label{eq:ec} \mathcal{E}_{\text{eff}}= \frac{1}{\Omega} \sqrt{\frac{\Omega}{\kappa}} \mathcal{E}_\mu( \kappa \cos \boldsymbol{k}_\mu r - 2 \Delta_{12} ), \end{align} where $\mathcal{E}_\mu = |\mathcal{E}_{\mu, i}|$, $ \kappa \equiv \Omega - \Delta_{13}$, and $\Omega$ is defined in Fig.~\ref{fig:1}. In Eq.~\eqref{eq:ec} we have chosen $\omega_\mu = \tfrac{1}{2}(\Delta_{13} - \Omega )$, so that $\mathcal{E}_{\text{eff}}$ is resonant with the {\sf a-b} transition. To ensure $\ket{\sf b}$ is prepared to high fidelity, $\Omega \gg \gamma$. Due to the divergence of the dipole-dipole interaction, this is naturally satisfied at qubit separations small compared to $\lambda_0$. For an NV center in diamond, the separation between qubits for the purposes of preparation is assumed to be $r = 50\text{nm}$, or $\xi_{12} = 0.5$. This is within the capabilities of present technology. In fact, there are proposals for $r \approx 1\text{nm}$ with the position of the qubit known to sub-nm accuracy~\cite{Me05a}. The effect of the laser taking into account the full eight-level Hamiltonian (assuming no photon emission) has been determined numerically [see Fig.~\ref{fig:2a}]. To quantify the success of the transfer {\sf a-b}, we use fidelity $(F)$ as a distance measure~\cite{Niel00}. In order to maintain high fidelity, the Rabi frequency $\mathcal{E}_\mu$ cannot be made arbitrarily large. This is because the coupling $\mathcal{E}_{\sf ad} \propto \text{e}^{\text{i} \Omega t}$. So, increasing $\mathcal{E}_\mu$ without altering $\Omega$ means the coupling {\sf a-d} will no longer be rapidly oscillating compared to the coupling {\sf a-b} [Eq.~\eqref{eq:ec}]. Note also that $\mathcal{E}_{\sf ac} \propto \text{e}^{\text{i} \Omega t}$, but since the magnitude of the coupling {\sf a-c} is much less than the magnitude of the coupling {\sf a-d}, {\sf a-c} is weakly coupled compared to {\sf a-d}. \begin{table}[bp] \begin{tabular}{lccc} & $F > 0.90$ & $F > 0.95$ & $F > 0.98$ \\ \colrule Rabi frequency, $\mathcal{E}_\mu$ & $\pm 20\%$ & $\pm 12.5\%$ & $\pm 6 \%$ \\ Detuning, $\omega_\mu$ & $\pm 5\tfrac{1}{2}\%$ & $\pm 4\%$ & $\pm 1\tfrac{1}{2}\%$ \\ Position variance, $v$ & & 0.08 & 0.005\\ \end{tabular} \caption{\label{tab:prep} Level of control for preparation of $\ket{\sf b}$ for exact values quoted in Fig.~\ref{fig:2a} with variance $v$. For $F > 0.9$, the overlap of the distribution of the position of two NV centers is large, causing their order to change. This renders a value for $v$ meaningless.} \end{table} We now focus our attention on the robustness of the preparation to variations in Rabi frequency, detuning, and separation (see Table~\ref{tab:prep}). We assume the Rabi frequency is equal to some value $\mathcal{E}$, rather than $\mathcal{E}_\mu = \gamma$. The fidelity of the preparation after time $t_\pi$---the time taken for population inversion---remains high, even for $\mathcal{E}/\mathcal{E}_\mu = \tfrac{3}{4}$. Thus, preparation is robust to small variations in $\mathcal{E}_\mu$. For detuning, we assume the laser frequency equals some value $\omega$, rather than $\omega_\mu$. For separation, we assume that in practice there will always be some uncontrollable variation in qubit separation, so the positions of the qubits are known only to a certain error. However, once they are positioned, they do not move. This is peculiar to NV centers in diamond and may not apply to other physical implementations of our proposal (e.g., atom traps). We assume that the probability distribution of the position of qubit $i$ is a Gaussian with mean zero and variance $v$ in units of $\lambda_0$ centered at $\boldsymbol{r}_i$. The position of the qubit is taken from this distribution using Monte Carlo techniques. Table~\ref{tab:prep} shows the limits on separation inaccuracy, averaged over $100$ different initial positions. If proposal~\cite{Me05a} is implemented, then typical variations of the NV centers will be $\sim$nm, which is $\sim 0.0015 \lambda_0$, the frequency of typical lasers can be controlled to within one part in $10^8$, and their amplitude varies by less than $0.25\%$ every $10$ seconds~\cite{newfocus}. So, in light of Table~\ref{tab:prep} high-fidelity preparation in NV centers, although certainly an experimental challenge, is possible with present technology. \begin{figure}[tbp] \begin{center} \subfigure[]{\label{fig:3a} \includegraphics[width=3.5cm,height=2.5cm]{fig3a.eps}} \subfigure[]{\label{fig:3b} \includegraphics[width=3.5cm,height=3.0cm]{fig3b.eps}} \end{center} \caption{(a) \label{fig:3} Population of the levels {\sf a,\ldots,h} with $\xi_{12} = 0.15$, $\mathcal{E}_\mu = 6 \gamma$, $\mathcal{E}_\nu = 15 \gamma$, $\omega_\delta = 170\gamma$, and $\alpha = \tfrac{\pi}{2}$. For these parameters, $F=0.986$ and the time taken for the transfer {\sf b-c}: $t_{\pi} = 9.271 \gamma^{-1}$. (b) Number of qubit rotations {\sf b-c} possible during time period $\tfrac{1}{2}(\gamma_{\sf b} + \gamma_{\sf c})^{-1}$ versus qubit separation with $F = 0.98$ and $\alpha = \tfrac{\pi}{2}$. } \end{figure} Before describing our method to rotate a DFS qubit, we explicitly include the probability of decay. The imaginary part of the eigenvalues of Eq.~\eqref{eq:hnh} give the linewidths of the eigenstates. The linewidth of $\ket{\sf b}$, $\gamma_{\sf b}$, decreases with decreasing qubit separation. Thus, we examine the number of population inversions possible in time period $\gamma_{\sf b}^{-1}$ with respect to physical qubit separation [see Fig.~\ref{fig:2b}]. This improves dramatically for small separations. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Logical qubit rotation} \label{sec:manip} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In order to rotate the encoded qubit, we use a bichromatic laser field. With the appropriate laser detunings this causes population between $\ket{\sf b}$ and $\ket{\sf c}$ to undergo coherent oscillations, without populating other eigenstates. At nonzero separation the DFSs are mixed by spontaneous emission, so in order to ignore this effect, we choose the lowest-energy states. The laser frequencies are chosen so that \begin{align} \omega_\mu = \omega_\delta \qquad \text{and} \qquad \omega_\nu = \tfrac{1}{2}(3\Delta_{13} - \Omega ) + \omega_\delta, \end{align} where $\omega_\mu$ ($\omega_\nu$) is resonant with transition {\sf b-e}({\sf c-e}). Similar to a two-photon Raman transition in an isolated three-level system, $\omega_\mu$ and $\omega_\nu$ are detuned from resonance by $\omega_\delta$. The effective couplings for the transitions {\sf b-e} and {\sf c-e} are \begin{align} \mathcal{E}_{\sf be} = \frac{1}{2 \Omega}\text{e}^{-\tfrac{\text{i}}{2} (\eta - 2 \omega_\delta)t}(\text{e}^{\tfrac{\text{i}}{2} \eta t} \mathcal{E}_\mu + \mathcal{E}_\nu)(\kappa - 8 \Delta_{12} \cos \boldsymbol{k} r) \end{align} and \begin{align} \mathcal{E}_{\sf ce} =& \sqrt{\frac{2\kappa}{\Omega}}\frac{\Delta_{12} \eta}{\Omega^2 - \Delta_{13}(\Delta_{13} + 4 \kappa)} \text{e}^{-\text{i}(\boldsymbol{k}r - \omega_\delta t)} \nonumber \\ &\times (1 - \text{e}^{2\text{i} \boldsymbol{k} r}) (\text{e}^{\tfrac{\text{i}}{2} \eta t} \mathcal{E}_\mu + \mathcal{E}_\nu) \end{align} for $\boldsymbol{k}$ orthogonal to the line joining the qubits with $\boldsymbol{k}_\mu = \boldsymbol{k}_\nu = \boldsymbol{k}$, and for $\eta \equiv \Omega - 3 \Delta_{13}$. Consider the three-level system {\sf b-c-e} in isolation, with the couplings calculated from $\widehat{H}_{\text{I}}$. After adiabatically eliminating $\ket{\sf e}$, the effective Rabi frequency between $\ket{\sf b}$ and $\ket{\sf c}$ in the collective basis and interaction picture is \begin{align} \label{eq:oef} \mathcal{E}_{\text{eff}} =& \frac{\text{e}^{-\text{i} (\boldsymbol{k} r + t\eta)}\kappa^{3/2}}{4\sqrt{2}\omega_\delta (\kappa \Delta_{13} - 2\Delta_{12}^2)^2 \Omega^{3/2}} (\text{e}^{2 \text{i} \boldsymbol{k} r} - 1) \Delta_{12} \nonumber \\ &\times \eta (\text{e}^{\tfrac{\text{i}}{2} t \eta} \mathcal{E}_\mu + \mathcal{E}_\nu) (\mathcal{E}_\mu + \text{e}^{\tfrac{\text{i}}{2} t \eta} \mathcal{E}_\nu)\nonumber \\ &\times (2\Delta_{12}^2 - \Delta_{13}\kappa - 2 \Delta_{12} \eta \cos \boldsymbol{k} r). \end{align} In order to calculate the detuning from resonance, $\omega_\delta$, the level shift caused by the interaction of the laser with the qubits was included. This was done numerically. We calculate the population (assuming no photon emission) of all eight states for the initial state $\ket{\sf b}$ [see Fig.~\ref{fig:3a}]. The effective Rabi frequency cannot be made arbitrarily large: this causes population excitation into $\ket{\sf e}$ and $\ket{\sf f}$, and, if the increase is large enough, transitions into $\ket{\sf h}$. If $\mathcal{E}_{\mu (\nu)} \gg \mathcal{E}_{\nu (\mu)}$, then the population will oscillate between {\sf b-e} ({\sf c-e}). In order for the adiabatic elimination of the upper state {\sf e} to remain valid, $\mathcal{E}_{\text{eff}} \ll \omega_\delta$. We examine the robustness of the logical qubit rotations to variations in Rabi frequency, detuning, and separation. For the Rabi frequency, we simultaneously vary $\mathcal{E}_\mu$ and $\mathcal{E}_\nu$ away from the exact values and examine $F$ at $t_\pi$, all of which are quoted in Fig.~\ref{fig:3a}. We examine the ratio $\mathcal{E}_\mu \mathcal{E}_\nu/90$ for $\mathcal{E}_\mu, \mathcal{E}_\nu$ chosen close to the values in Fig.~\ref{fig:3a} (see Table~\ref{tab:rot}). For high-fidelity rotations, the level of control of the Rabi frequency is high, but due to the small separation, greater tolerance for the detuning is permitted. Regarding NV centers, high-fidelity rotations are possible with sub-nm position control~\cite{Me05a}. Note that this control assumes that once the NV centers are located, their positions cannot be determined, and so the frequencies of the control fields cannot be adjusted accordingly. \begin{table}[t] \begin{tabular}{lccc} & $F > 0.90$ & $F > 0.95$ & $F > 0.98$ \\ \colrule Rabi frequency, $\mathcal{E}_\mu \mathcal{E}_\nu/90$ & $\pm 12\%$ & $\pm 8\%$ & $\pm 1 \%$ \\ Detuning, $\omega_\delta$ & $\pm 28\%$ & $\pm 18\%$ & $\pm 6\%$ \\ Position variance, $v$ & $1.6 \times 10^{-3}$ & $0.8 \times 10^{-3}$ & $6 \times 10^{-6}$ \\ \end{tabular} \caption{\label{tab:rot} Level of control for oscillations between {\sf b-c} for exact values quoted in Fig.~\ref{fig:3a} with variance $v$.} \end{table} The number of rotations possible in a single spontaneous emission lifetime enables us to see the gains obtained from using DFS encoding. For the linewidth of the qubit we use $\tfrac{1}{2}(\gamma_{\sf b} + \gamma_{\sf c})$, which is the average linewidth of the two states. Figure~\ref{fig:3b} shows the number of qubit rotations per spontaneous emission lifetime. As the separation increases, the number of rotations decreases. This is partly due to the increase in linewidth with increasing separation, but mainly due to $\Delta_{\pm}$ decreasing in size. This means the three-level system {\sf b-c-e} can no longer be treated in isolation. In fact, \begin{align} \lim_{\xi_{ij}\to \infty} \mathcal{E}_{\text{eff}} = 0. \end{align} At separations much larger than $\xi = 0.2$, or $20\text{nm}$, the time taken for the rotation is above that taken for decay from {\sf b} and {\sf c}, negating the benefits of the proposed DFS encoding for quantum-information processing. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{State read out} \label{sec:rout} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% This is performed in a similar way to preparation. We exploit the splitting of the eigenbasis in order to isolate a one-photon transition that will fluoresce if populated. For this purpose, a single laser field is required [see Eq.~\eqref{eq:slf}]. Similar to preparation, the detuning of the laser field is chosen to be equal to $\omega_\mu= \tfrac{1}{2}(3 \Delta_{13} + \Omega)$. So, the (resonant) coupling between states $\ket{\sf c}$ and $\ket{\sf g}$ is: \begin{align} \mathcal{E}_{\sf cg} = \frac{\text{i}}{\sqrt{2}} \sqrt{1 + \frac{\Delta_{13}}{\Omega}} \frac{\Delta_{12} (\Omega + 3 \Delta_{13})\mathcal{E}_\mu \sin \boldsymbol{k} r }{2 \Delta_{12}^2 + \Delta_{13} (\Omega + \Delta_{13})}. \end{align} If the qubit is in $\ket{1}_L$, then fluorescence will be detected, if not, then the system will remain dark. The decay width of state $\ket{\sf g}$ is \begin{align} \label{eq:glw} \gamma_{\sf g} = \frac{1}{2}\left( 4\gamma + \gamma_{13} + \sqrt{8 \gamma^2_{12} + \gamma^2_{13}}\right). \end{align} This is superradiant with an upper bound of $4 \gamma$. In order to read out $\ket{0}_L$, the frequency of the laser field is chosen to be resonant with the transition {\sf b-g}, i.e., $\omega_\mu=\Omega$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Two-qubit operations and cluster-state preparation} \label{sec:tqu} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In order to perform an arbitrary sequence of quantum logic operations two-qubit entangling operations are required~\cite{Niel00}. Here, two methods of performing these operations are described. The first is a natural extension of Sec.~\ref{sec:manip}, and the second applies the technique described in Ref.~\cite{Ch05} to a collection of systems of three closely spaced physical qubits in order to prepare a cluster state. \begin{figure}[t] \begin{center} \includegraphics[width=5.5cm,height=2.5cm]{fig4.eps} \end{center} \caption{\label{fig:cphase} CPHASE gate detailed in Ref.~\cite{Ch05} between atom $A$ and atom $B$. The $\tfrac{1}{4}$-wave plates are labeled with $\tfrac{\lambda}{4}$, and the polarization detector with $D$. A left-circularly polarized photon $\ket{\text L}$ is input at the left, and the subsequent polarization is measured at $D$.} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Two-qubit operations} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% The smallest number of physical qubits that supports two decoherence-free qubits is four. The decoherence-free qubits are encoded into a decoherence-free subsystem and a decoherence-free subspace. Using the Dicke decomposition, the decoherence-free subspace in four qubits is~\cite{Kem01} \begin{align} \ket{0}_L = &\frac{1}{2}(\ket{01}-\ket{10})(\ket{01}-\ket{10}), \\ \ket{1}_L = &\frac{1}{\sqrt{12}}(2\ket{0011} + 2\ket{1100} -\ket{0101} \nonumber \\ &-\ket{1010} -\ket{0110} -\ket{1001}). \end{align} The Hilbert space is decomposed into irreducible representations: $\mathcal{H}_1 \oplus \mathcal{H}_1 \oplus \mathcal{H}_3 \oplus \mathcal{H}_3 \oplus \mathcal{H}_3 \oplus \mathcal{H}_5$, where the subscript labels the dimension. The decoherence-free subspace qubit is encoded across $\mathcal{H}_1 \oplus \mathcal{H}_1$, and the decoherence-free subsystem qubit across $\mathcal{H}_3 \oplus \mathcal{H}_3$. The controlled-phase (CPHASE) gate, $\{\ket{00}_L, \ket{01}_L, \ket{10}_L, \ket{11}_L \}$ $\to$ $\{ \ket{00}_L, \ket{01}_L, \ket{10}_L, -\ket{11}_L \}$, is an entangling operation. So that two qubits will yield two logical bits of information, the logical states are defined in pairs: $\ket{\sf b} \equiv \ket{01}_L$, $\ket{\sf c} \equiv \ket{00}_L$, $\ket{\sf f} \equiv \ket{11}_L$, $\ket{\sf g} \equiv \ket{10}_L$, where the Hilbert space is labeled {\sf a,b, \ldots,o,p} in order of increasing energy. Using this labeling, a CPHASE operation is performed using a $2\pi$ pulse, off resonant with transition {\sf f-l}, that produces a phase shift of $-1$. Arbitrary one-qubit operations are performed on the first qubit by resonantly coupling {\sf b-f} and {\sf c-g} simultaneously, and on the second qubit by rotating {\sf b-c} and {\sf f-g} using collective two-photon Raman transitions as described in Sec.~\ref{sec:manip}. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection{Cluster-state preparation} \label{sec:cluster} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[t] \begin{center} \includegraphics[width=5.5cm,height=4.cm]{fig5.eps} \end{center} \caption{\label{fig:dfcavity} The decoherence-free qubit is trapped in a one-sided optical cavity. The transition {\sf c-g} is coupled resonantly to the right circularly polarized mode of the cavity with coupling constant $g$. The cavity photon is either transmitted through the cavity mirror with rate $\kappa_c$ or lost with rate $\kappa_l$. $b_{in}(t)$ and $b_{out}(t)$ denote the input and output field operators, respectively.} \end{figure} Although the previous method supports two-qubit operations, it is not conducive to QI processing in large systems: the entangling gate works only with isolated systems of neighboring qubits. One method that enables arbitrary quantum-logic operations is one-way computation~\cite{Bri01Rau01}. In order to generate the cluster states required for this, we apply the method proposed in Cho and Lee~\cite{Ch05} to a collection of spatially separated arrays of three dipole-coupled qubits that are situated in cavities. In Ref.~\cite{Ch05}, the (atomic) qubit consists of the lower levels of a three-level atom (3LA), and is situated in a one-sided optical cavity. The right circularly polarized mode of the cavity photon resonantly couples one of the logical states (e.g., $\ket{1}$) to the upper state of the 3LA, but not the other. The phase of a right circularly polarized photon on exiting the cavity is unchanged if the qubit is in $\ket{1}$, otherwise the photon, whether right or left circularly polarized, acquires a $\pi$-phase shift. A CPHASE gate between two separated atomic qubits can be realized using the setup shown in Fig.~\ref{fig:cphase}. This system can be realized here by placing the logical qubit inside a cavity (see Fig.~\ref{fig:dfcavity}). The encoded qubit is placed in a cavity on resonance with $\frac{1}{2}\left(3 \Delta_{13} + \Omega \right)$ so only the right circularly polarized cavity photon interacts with the qubit in state $\ket{\sf c}$. If the qubit is in $\ket{\sf b}$, the cavity photon acquires a $\pi$-phase shift, otherwise it does not. Then, two-qubit operations are performed in the same manner as described in Ref.~\cite{Ch05}. However, our setup has a number of further requirements. First, the atom-cavity coupling rate has to be fast compared to the decay time scale of the DF state $\ket{\sf c}$. Second, $\ket{\sf g}$ is superradiant [Eq.~\eqref{eq:glw}], so the atom-cavity coupling rate must be fast compared to $\gamma_{\sf g}$. Fortunately, the upper bound of $\gamma_{\sf g}$ is $4\gamma$, which is small compared to $\Delta_\pm$. Third, as well as the probability of logical qubit decay, if there is a decay from {\sf g}, the most probable decay channel is {\sf g-d} not {\sf g-c}, so any decay implies information loss. Cluster states can be generated as follows~\cite{Ch05}. The $1$D cluster state of $N$ qubits can be written as \begin{align} \label{eq:clus} \ket{\Psi_{N}}& = \frac{1}{\sqrt{2}}\ket{\phi_0}_{N-3} \ket{0}_{N-2} \left(\ket{0}_{N-1} \ket{+}_N + \ket{1}_{N-1} \ket{-}_N \right)\nonumber \\ +& \frac{1}{\sqrt{2}}\ket{\phi_1}_{N-3} \ket{1}_{N-2} \left(\ket{0}_{N-1} \ket{+}_N - \ket{1}_{N-1} \ket{-}_N \right)\end{align} for $\ket{\pm} = \tfrac{1}{\sqrt{2}}(\ket{0} \pm \ket{1})$, where the subscript labels the qubit, and the terms for the $(N-3)$ qubits are denoted by $\ket{\phi_i}$. The state $\ket{\Psi_{N+1}}$ can be generated by attaching a qubit in $\ket{+}$ to $\ket{\Psi_{N}}$ by performing a CPHASE operation. If this operation fails, state~\eqref{eq:clus} becomes a mixed state. However, $\ket{\Psi_{N - 2}}$ can be recovered from this state by measuring the $(N-1)$th qubit in the computational basis and performing an operation on the $(N-2)$nd qubit dependent on the measurement result. The average number of qubits attached to the cluster state by $m$ CPHASE operations is $(3P - 2)m$, which grows if $P > 2/3$. So, cluster states consisting of DF logical qubits inside separate cavities can be prepared to high-fidelity using linear optics. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Conclusion} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% We have proposed a general method to prepare a maximally entangled DFS state in a linear array of dipole-coupled qubits that, although relevant to any system of dipole-coupled qubits, in the light of recent experimental progress, is directly applicable to a system of NV centers in diamond. The preparation is possible to high-fidelity, and can be performed quickly relative to the linewidth of the entangled state. We showed how to rotate a logical qubit to high-fidelity within a DFS without leakage, and quickly relative to the combined decay of the logical states. Similar to the preparation, we showed how to read out a logical state with high-fidelity. We then described two methods to perform two-qubit operations, one of which enables cluster-state preparation in a system of spatially separated DF dipole-coupled qubits. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Acknowledgments} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% We especially thank Karl-Peter Marzlin for many helpful discussions. We also thank Jim Cresser, Barry Sanders, and Jason Twamley for comments on the manuscript. This work was supported by Macquarie University. \begin{thebibliography}{99} \bibitem{Pal96Duan97Zan97aLidar98Knill00}{G.M. Palma, K.-A. Suominen, and A. K. Ekert}, {Proc. Roy. Soc. London Ser. A}, \textbf{452}, 567 (1996); {L.-M Duan and G.-C. Guo}, \prl \textbf{79}, 1953 (1997); {P. Zanardi and M. Rasetti}, {\it ibid}. \textbf{79}, 3306 (1997); {D.A. Lidar, I. L. Chuang, and K. B. Whaley}, {\it ibid}. \textbf{81}, 2594 (1998); {E. Knill, R. Laflamme, and L. Viola}, {\it ibid}. \textbf{84}, 2525 (2000). \bibitem{Be00Fe02Zan98}{A. Beige, D. Braun, B. Tregenna, and P. L. Knight}, \prl \textbf{85}, 1762 (2000); {M. Feng and X. Wang}, \pra \textbf{65}, 044304 (2002); {P. Zanardi and F. Rossi}, \prl \textbf{81}, 4752 (1998). \bibitem{Pet02}{D. Petrosyan and G. Kurizki}, \prl \textbf{89}, 207902 (2002). \bibitem{Kem01}{J. Kempe, D. Bacon, D. A. Lidar, and K. B. Whaley}, \pra \textbf{63}, 042307 (2001). \bibitem{Kiel01}{D. Kielpinski \emph{et al}.}, Science \textbf{291}, 1013 (2001). \bibitem{Kw00}{P. G. Kwiat, A. J. Berglund, J. B. Altepeter, and A. G. White}, Science \textbf{290}, 498 (2000). \bibitem{Bou04}{M. Bourennane, M. Eibl, S. Gaertner, C. Kurtsiefer, A. Cabello, and H. Weinfurter}, \prl \textbf{92}, 107901 (2004). \bibitem{Moh03}{M. Mohseni, J. S. Lundeen, K. J. Resch, and A. M. Steinberg}, \prl \textbf{91}, 187903 (2003). \bibitem{Lo01} L. Viola {\it et al}., Science \textbf{293}, 2059 (2001). \bibitem{Da94} G. Davies, {\it Properties and growth of diamond} (IEE/INSPEC, London, 1994), Vol. \textbf{9}. \bibitem{Je04} F. Jelezko \emph{et al}., \prl \textbf{93}, 130501 (2004). \bibitem{Ho05} M. Howard \emph{et al}., New J. Phys. \textbf{8}, 33 (2006). \bibitem{Me05Ra05} J. Meijer \emph{et al}., {Appl. Phys. Lett.} \textbf{87}, 261909 (2005); J.R. Rabeau \emph{et al}., {\it ibid}. (to be published). \bibitem{Ni03Ni03a} A.P. Nizovtsev \emph{et al}., {Opt. Spectrosc.} \textbf{94}, 848 (2003); {Physica B} \textbf{340}, 106 (2003). \bibitem{Dr99} A. Drabenstedt \emph{et al}., \prb \textbf{60}, 11503 (1999). \bibitem{San06} C. Santori \emph{et al}., Opt. Express \textbf{14}, 7986 (2006). \bibitem{Mil74}{P. W. Milonni and P. L. Knight}, \pra \textbf{10}, 1096 (1974). \bibitem{Car93}{H.J. Carmichael}, {\em{An Open Systems Approach to Quantum Optics}} (Springer-Verlag, Berlin, 1993). \bibitem{Dicke} R.H. Dicke, Phys. Rev. \textbf{93}, 99 (1954). \bibitem{Clem03}{J. P. Clemens, L. Horvath, B. C. Sanders, and H. J. Carmichael}, \pra \textbf{68}, 023809 (2003). \bibitem{Bela69Lehm70iLehm70iiArg70} A.A. Belavkin \emph{et al}., {Sov. Phys. JETP} \textbf{56}, 264 (1969); G.S. Agarwal, \pra{\textbf{ 2}}, 2038 (1970); R.H. Lehmberg, {\it ibid}. {\textbf{ 2}}, 883 (1970); {\textbf{ 2}} 889 (1970). \bibitem{Lind}{G. Lindblad}, Commun. Math. Phys. \textbf{48}, 119 (1976) \bibitem{Me05a} J. Meijer \emph{et al}., Appl. Phys. A \textbf{83}, 321 (2006). \bibitem{Niel00}{M.A. Nielsen and I.L. Chuang}, {\it Quantum Computation and Quantum Information} (Cambridge University Press, Cambridge, England, 2000). \bibitem{newfocus} See, e.g., the specifications published by the manufacturers New Focus at http://www.newfocus.com. \bibitem{Bri01Rau01}{H.J. Briegel and R. Raussendorf}, \prl \textbf{86}, 910 (2001); {R. Raussendorf and H.J. Briegel}, {\it ibid}. \textbf{86}, 5188 (2001). \bibitem{Ch05} J. Cho and H.-W. Lee, \prl \textbf{95}, 160501 (2005). \end{thebibliography} \end{document} }
\caption{Time evolution of the $z$-component of the angular momentum expectation value for $\varepsilon = 0$, $v =1$ and $c=0.5/(N+1)$. The many particle dynamics ($\textcolor{green}{\bullet}$ ) for an initial coherent state $\left|N\right>$ located at the north pole is compared with the mean-field trajectory ($\textcolor{blue}{\bullet}$) for a starting vector $\boldsymbol s/s = \boldsymbol x = (0,0,1)$ (the lines are plotted to guide the eye); left: $N = 50$ particles, right: $N = 100$ particles.} \label{Figure:mean-Nteil_c_0-5_N_33} \end{figure} Figure \ref{Figure:mean-Nteil_c_0-5_N_33} shows a comparison of the mean-field and many particle dynamics. Shown is the $z$-component of the angular momentum expectation value (i.e.~the population imbalance of the two wells) for a symmetric system ($\varepsilon = 0$) with $v=1$ and a weak interaction strength $c = 0.5/(N+1)$, where the initial state is a coherent state $\left|N\right>$ located at the north pole, or $\boldsymbol s/s = (0,0,1)$ in the mean-field approximation. Both oscillate with frequency $\omega=1$ (compare \eqref{omega}) and the envelope of the many particle expectation values decay with a width $\sim \sqrt{N+1}$. Note, however, that the deviation between the mean-field and many particle dynamics is due to the fact that the mean-field description is based on a single trajectory and can be cured to some extent by propagating an ensemble approximating initially the coherent state distribution of the many particle system \cite{07meanf}. Our interest, however, is not focused on the regular regime, but on the mixed regular-chaotic regime with the two prominent regular islands centered at the self-trapping states $\boldsymbol s_\pm = \boldsymbol s(\vartheta_\pm,\varphi_\pm)$. A mean-field trajectory started in one of these states will stay there forever. The full quantum system however, shows a much richer behavior here. Starting the propagation in one of the corresponding coherent states $\left|\pm\right> = \left|\vartheta_\pm,\varphi_\pm\right>$ the system exhibits dynamical tunneling to the other island, i.e.~the state $\left|\mp\right>$. Such dynamical tunneling processes are known for many systems \cite{Grif98,Bayf99}, e.g., a particle in periodically driven double-well potential \cite{Lin90, Lin92, Bavl93} or a periodically driven rotor \cite{95tun,Bonc98, Aver02}. Such a system can be realized with a BEC \cite{Hens00, Hens01a, Hens01b, Stec01, Salm02, Osov05} allowing direct experimental observation of the tunneling. \begin{figure} \centering \includegraphics[scale=1]{eps/tunnel_c_2_fixh.eps} \includegraphics[scale=1]{eps/tunnel_c_2_N_10_t_40h.eps} \includegraphics[scale=1]{eps/tunnel_c_2_N_20_t_1000h.eps} \includegraphics[scale=1]{eps/tunnel_c_2_N_100_t_1000h.eps} \caption{a) Mean-field dynamics for the parameters $\varepsilon = 0$, $v=1$, $\tau = 1$ and $c=2/(N+1)$, the fixed points $\boldsymbol s_\pm$ are marked with red dots. b) Propagation of many-particle state $\left|-\right>$ located on the southern fixed point for $N=10$ particles; the dynamical tunneling to the northern fixed point lasts $40$ periods here; c) as before, but $N = 20$ particles; tunneling during $1000$ periods d) as in c), but $N=100$ particles, dynamical tunneling cannot be observed here.} \label{Figure:tunnel_c_2_fix} \end{figure} It should be emphasized, that here the dynamical tunneling between both of the island states is a coherent many particle process and may not be confused with the single particle tunneling which causes the Rabi oscillations. Nevertheless it corresponds to an actual tunneling of the condensate between the potential wells, since the northern state is localized almost in the first well, whereas the southern state occupies mostly the second well. In figure \ref{Figure:tunnel_c_2_fix} the tunneling process starting from the fixed point on the southern hemisphere in the state $\left|\psi(0)\right>= \left|-\right>$ is shown for several parameter sets. An interaction strength of $c = 2/(N+1)$ is chosen to ensure the two regular islands to be not too small and well separated. One observes dynamical tunneling through the Bloch sphere. The normalized expectation value $\left<\boldsymbol L\right>/\ell$ spirals its way on a tube encircling the straight line connecting the two island states to the other side of the sphere. The radius of the tunneling tube is determined by the uncertainty of the direction of the angular momentum vector in a coherent state, which is proportional to $1/N$. The more particles one puts into the system, the tighter is the spiral around the connection line of the fixed points and the longer gets the tunneling time. However, under some conditions the tunneling time can also decrease with a growing number of particles, leading to a strongly enhanced population transfer, as will be explained in section \ref{subsecparameter}. But let us first have a closer look at the tunneling process. The squared absolute values $\left|\left<\psi(t)\big|\pm\right>\right|^2$ of the projections of $\left|\psi(t)\right>$ on the coherent island states and the orthogonal subspace for $\left|\psi(0)\right>=\left|-\right>$ for case c) of figure \ref{Figure:tunnel_c_2_fix} with $N = 20$ is shown in figure \ref{Figure:inselproj_N_20_c_2} as a function of time. \begin{figure} \centering \includegraphics[scale=1]{eps/inselproj_N_20_c_2.eps} \caption{Squared absolute values of the projections of $\left|\psi(t)\right>$ on the coherent states $\left|\pm\right>$ during propagation over $1000$ periods for $N=20$ particles, $\varepsilon = 0$, $v=1$, $\tau=1$ and $c=2/(N+1)$; $\textcolor{blue}{-}$ $\left|\left<-\big|\psi(t)\right>\right|^2$, $\textcolor{green}{-}$ $\left|\left<+\big|\psi(t)\right>\right|^2$ and $\textcolor{red}{-}$ orthogonal subspace.} \label{Figure:inselproj_N_20_c_2} \end{figure} After $1000$ periods, the population has been transferred almost completely to the state $\left|+\right>$ on the northern hemisphere. For all particle numbers $N$ considered in figure \ref{Figure:tunnel_c_2_fix}, the population $\left|\left<\kappa\big|-\right>\right|^2$ of the southern coherent island state on the Floquet states $\left|\kappa\right>$ is significantly large only for two of the $(N+1)$ Floquet states which are denoted by $\left|\kappa_\pm\right>$ with corresponding quasi-energies $\epsilon_\pm$. As indicated by the $\pm$-sign, they belong to different symmetry classes and the island states can be reconstructed up to 94\,\% by the linear combinations \begin{equation}\label{k4_linkomb} \left|\psi_\pm\right> = \frac{1}{\sqrt{2}}\left(\left|\kappa_+\right> \mp \left|\kappa_-\right>\right) \end{equation} which hence are a good approximation for the coherent states $\left|\pm\right>\approx \left|\psi_\pm\right>$. If one chooses $\left|\psi_-\right>$ as a starting vector, $\left|\psi(t)\right>$ always stays a linear combination of the $\left|\kappa_\pm\right>$ during propagation. Remembering equation \eqref{k3_Feigenzu} for the decomposition of the Floquet states, the propagation over $n$ kick periods yields \begin{eqnarray} \left|\psi(n\tau)\right> &=& \frac{1}{\sqrt{2}}{\rm e}^{-{\rm i}n\epsilon_+\tau }\left(\left|\kappa_+\right> + {\rm e}^{-{\rm i}n(\epsilon_- - \epsilon_+)\tau}\left|\kappa_-\right>\right)\\ &=& \frac{1}{2}{\rm e}^{-{\rm i}n\epsilon_+\tau }\left( \left(1 - {\rm e}^{-{\rm i}\phi_n} \right)\left|+\right> + \left(1 + {\rm e}^{-{\rm i}\phi_n} \right)\left|-\right> \right), \end{eqnarray} with $\phi_n = n(\epsilon_- - \epsilon_+)\tau$. The expectation value of the angular momentum varies as \begin{equation}\label{ellipse} \left<\boldsymbol L (n\tau) \right> = \frac{1-\cos{\phi_n}}{2}\boldsymbol L_+ + \frac{1+\cos{\phi_n}}{2}\boldsymbol L_- + \sin{\phi_n}{\rm Im}{\boldsymbol L_{-+}} \end{equation} with $ \boldsymbol L_\pm = \left<\pm\right|\boldsymbol L \left|\pm\right>$, $ \boldsymbol L_{-+} = \left<-\right|\boldsymbol L \left|+\right>$, where in the present cases ${\rm Im}\boldsymbol L_{-+}$ is zero or very small so that the ellipse \eqref{ellipse} degenerates to a straight line connecting the island centers at $\boldsymbol L_-$ and $\boldsymbol L_+$. Thus the spiraling on a tube around this line is due to the other contributing Floquet states. We conclude, that the subsystem consisting of the two island states $\left|\pm\right>$ can be described in a good approximation by a two state system of the Floquet states $\left|\kappa_\pm\right>$. Thus, as any two level system, the system exhibits periodic tunneling with period \begin{equation}\label{k4_tunnelzeit} T_{\rm tunnel} = \frac{2\pi}{\epsilon_- - \epsilon_+} = \frac{2\pi}{\Delta\epsilon}. \end{equation} These considerations are entirely equivalent to the tunneling process through a symmetric potential barrier, where the quasi-energies adopt the function of the energies of time independent systems. Thus the tunneling period is solely ruled by the quasi-energy splitting $\Delta\epsilon$. One should keep in mind, that despite the similarities in the formalism this is no single particle, but a coherent many particle process. Still the dynamical tunneling between the two island states corresponds to an actual tunneling of the BEC from one potential well to the other. If the system is localized in the southern island $\left|-\right>$, the expectation value of $L_z$ is negative and the condensate is located mainly in the second well; in the case $c = 2/(N+1)$ over $80\,\%$ of the condensate is trapped there. The opposite is true for the northern island $\left|+\right>$. To give an impression of the localization of the Floquet states, one can compute quantum phase space distributions, e.g.~the Husimi distributions \cite{Husi40, Hine05} \begin{equation} Q(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi\right>\right|^2. \end{equation} via the $SU(2)$ coherent states $\left|\vartheta,\varphi\right>$. In figure \ref{Figure:husimi-insel}, the Husimi distribution for the linear combinations $\left|\psi_\pm\right>$ of the two Floquet tunneling states are shown. \begin{figure} \centering \includegraphics[scale=1]{./eps/hus_minus_c_2_N_30.eps}\includegraphics[scale=1]{./eps/hus_plus_c_2_N_30.eps} \caption{Husimi distributions of the island states $\left|\psi_\pm\right>$; left: $Q_-(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi_-\right>\right|^2$, right: $Q_+(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi_+\right>\right|^2$ for $N=20$ particles, $\varepsilon = 0$, $v=1$, $\tau=1$ and $c=2/(N+1)$.} \label{Figure:husimi-insel} \end{figure} The sharp localization in one of the two regular islands is evident here. \subsection{Parameter dependencies of the tunneling time}\label{subsecparameter} In general the dynamics of the island states, however, is more complicated and shows special features. Tunneling can in fact be extremely stimulated or fully suppressed as we will see in the following. At first we examine the tunneling for different numbers of particles. In figure \ref{Figure:tunnelzeit_c_2_N_0_150} the tunneling period $T_{\rm tunnel}$ for $c=2/(N+1)$ calculated with the help of \eqref{k4_tunnelzeit} is plotted over $N$; please note the logarithmic scaling of the $T$-axis. This is in excellent agreement with the tunneling time directly determined from the time propagation, plotted in red for comparison. \begin{figure} \centering \includegraphics[scale=1]{eps/tunnelzeit_c_2_N_0_150.eps} % tunnelzeit_c_2_N_0_150.eps: 300dpi, width=3.84cm, height=1.45cm, bb= 70 335 523 506 \caption{Tunneling period $T_{\rm tunnel}$ calculated from $\Delta\epsilon$ according to \eqref{k4_tunnelzeit} as a function of $N$ for $\varepsilon = 0$, $v = 1$, $\tau=1$ and $c = 2/(N+1)$; $\textcolor{red}{\bullet}$ $T_{\rm tunnel}$ determined from numerical propagation.} \label{Figure:tunnelzeit_c_2_N_0_150} \end{figure} For small $N$ the tunneling period increases exponentially. For larger $N$ new structures appear, where the tunneling time decreases rapidly about several orders of magnitude. These ``resonances'' have been observed for other systems e.g.~the anharmonic driven oscillator \cite{Lin90, Shin94, Bonc98}. Quite often this phenomenon is called ``chaos assisted tunneling'' \cite{Lin90, Lin92, Plat92, Uter94, Haen94, Grif98} or ``resonance assisted tunneling'' \cite{brod01,Elts05} and can be explained as an avoided crossing of the tunneling doublet $\epsilon_\pm$ with some third level $\epsilon_c$ which increases the splitting $\Delta \epsilon$. In such regions one actually has to deal with a three state tunneling mechanism. Nevertheless the tunneling period can still be calculated via \eqref{k4_tunnelzeit} quite accurately. We illustrate this by means of the resonance at $N\approx33$ where $T_{\rm tunnel}$ changes by two orders of magnitude. The Husimi distributions of the three Floquet states $\left|\kappa_\pm\right>$ and $\left|\kappa_c\right>$ involved in the tunneling for $N = 33$ are shown in figure \ref{Figure:hus_N_33_c_2}. \begin{figure} \centering \includegraphics[scale=1]{eps/hus_vgl_c_2.eps} \includegraphics[scale=1]{eps/hus_N_33_c_2_K_10.eps} \includegraphics[scale=1]{eps/hus_N_33_c_2_K_11.eps} \includegraphics[scale=1]{eps/hus_N_33_c_2_K_9.eps} \caption{The three tunneling states for $N=33$, $\varepsilon = 0$, $v=1$, $\tau=1$ and $c=2/(N+1)$: a) mean-field phase space , b) Husimi distribution $\left|\left<\vartheta,\varphi\big|\kappa_+\right>\right|^2$, c) $\left|\left<\vartheta,\varphi\big|\kappa_-\right>\right|^2$, d) $\left|\left<\vartheta,\varphi\big|\kappa_c\right>\right|^2$.} \label{Figure:hus_N_33_c_2} \end{figure} The third state $\left|\kappa_c\right>$ is mostly localized in the chaotic regions of the classical phase space and may therefore be denoted as ``chaotic'' state. The overlap of this state with the island states is actually small, $\left|\left<\kappa_c\big|\pm\right>\right|^2 \approx 0.06$. Nevertheless it changes the behavior of the system dramatically. To visualize the tunneling process, it is sufficient to focus on the $z$-component during propagation. In figure \ref{Figure:sz_N_30u33_c_2_t_10000} the $z$-component is shown for $N = 30$ and $N = 33$. \begin{figure} \centering \includegraphics[scale=1]{eps/sz_N_30u33_c_2_t_10000.eps} \caption{Population imbalance $\left<L_z\right>/\ell$ over $10000$ kick periods for $\varepsilon =0$, $v=1$ and $c = 2/(N+1)$; $\textcolor{green}{-}$ $N=30$ particles, $\textcolor{blue}{-}$ $N=33$ particles.} \label{Figure:sz_N_30u33_c_2_t_10000} \end{figure} The great difference in the tunneling period is obvious here. The system with $N = 30$ particles shows perfect self-trapping up to $10^4$ kick periods, whereas the system with just three more particles exhibits strong tunneling with $T_{\rm tunnel} \approx 4000\tau$. Thus, the small variation of the number of particles obviously changes the entire tunneling process from a two state to a faster three state mechanism. The interplay with the third ``chaotic'' state motivates the name ``chaos assisted tunneling'' (CAT) for this decrease of the tunneling period. If one takes a closer look on the tunneling oscillations of this system the small modulations attract attention. This can easily be understood as tunneling into the third state $\left|\kappa_c\right>$ which is superimposed. The period $T_c$ of these modulations can be calculated by the means of \eqref{k4_tunnelzeit} with the quasi-energy splitting $\Delta\epsilon_c$ between the third level and the tunneling doublet. In the present case this yields $T_c = 551$ in perfect agreement with the numerical propagation. To achieve a better understanding of the behavior of the quasi-energies, especially crossing or avoided crossing scenarios in regions where the levels really get close to each other, one needs to examine their dependence on continuous variables, e.g., the kick strength $c$. Of special interest is the role of the third chaotic state. One has to keep in mind, however, that the mean-field phase space structure changes by varying $c$; especially both of the regular islands are modified. At first we take a look at the tunneling time $T_{\rm tunnel}$ as a function of $c$ where we leave the number of particles $N=30$ fixed. \begin{figure} \centering \includegraphics[scale=1]{eps/tunnel_c_1_1-3_1_N_30.eps} \caption{Tunneling time $T_{\rm tunnel}$ calculated according to \eqref{k4_tunnelzeit} with the help of $\Delta\epsilon$ as a function of $c$ for a fixed number of $N = 30$ particles, $\varepsilon = 0$, $v=1$ and $\tau=1$; CDT: coherent destruction of tunneling divergences, CAT: chaos assisted tunneling resonances, in the regions of the small gaps the two state approximation breaks down.} \label{Figure:tunnel_c_1-1_3-1_N_30} \end{figure} This plot, shown in figure \ref{Figure:tunnel_c_1-1_3-1_N_30}, provides similar features as seen for the $N$-dependence in figure \ref{Figure:tunnelzeit_c_2_N_0_150}. For higher values of $c$ there exist the same kind of CAT resonances as seen in the $N$ dependence above, where the tunneling period decreases by several orders of magnitude. They are due to avoided crossing scenarios of the quasi-energy levels, which shall be examined in more detail in the following. \begin{figure} \centering \includegraphics[scale=1]{eps/kreuzung00.eps} \caption{Quasi-energies $\epsilon_\kappa$ for $N=30$ particles, $\varepsilon = 0$, $v=1$ and $\tau=1$ as a function of $c$; $\textcolor{red}{\bullet}$ tunneling doublet $\epsilon_\pm$; (CDT): real crossing of $\epsilon_\pm$ inducing CDT; (CAT): avoided crossing with a chaotic level inducing CAT, see text for details.} \label{Figure:kreuzung00} \end{figure} In figure \ref{Figure:kreuzung00} we show the quasi-energies over the interaction strength $c$ for $N=30$ particles (note the cylindrical topology; due to the definition of the quasi-energies $-\pi$ and $\pi$ have to be identified on the $\epsilon$-axis). Let us focus on the prominent resonance at $c(N+1) \approx 2.35$ corresponding to the region marked by CAT in the figure. A magnification of this region is shown in figure \ref{Figure:kreuzung1}. \begin{figure} \centering \includegraphics[scale=1]{eps/kreuzung1.eps} \caption{Blow-up of the CAT region of figure \ref{Figure:kreuzung00}; the symmetry of the quasi-energy levels are denoted by $(\pm)$; one observes an avoided crossing of the quasi-energy levels with $(+)$-symmetry; $\textcolor{red}{\bullet}$ tunneling doublet $\epsilon_\pm$.} \label{Figure:kreuzung1} \end{figure} Here one of the levels of the doublet undergoes an avoided crossing with a chaotic third level $\epsilon_c$ (since they belong to the same symmetry class). The splitting of the avoided crossing is in the same order of magnitude as the splitting of the tunneling doublet. Hence, in an intermediate region the three states $\left|\kappa_\pm\right>$ and $\left|\kappa_c\right>$ all take part in a three state tunneling mechanism. The simple two state model breaks down here and the tunneling time cannot be evaluated according to \eqref{k4_tunnelzeit} (note the small gaps in the graph in figure \ref{Figure:tunnel_c_1-1_3-1_N_30}). Due to the strongly avoided crossing the splitting of the tunneling doublet increases, when the third state approaches, which causes a significant decrease of the tunneling time. An illustration of this stimulated tunneling process, CAT, can be found in \cite{Toms94,Kohl98}. Comparing figure \ref{Figure:tunnel_c_1-1_3-1_N_30} and \ref{Figure:kreuzung00} it becomes obvious that apart from these resonant avoided crossing events there is a large number of avoided crossings which have no significant influence on the tunneling period. The avoided crossing between the third level and one of the doublet levels has only a very small splitting in these cases, which has no influence on the tunneling period, since the quasi-energy doublet stays almost undistorted. Despite the CAT resonance there is another structure in figure \ref{Figure:tunnel_c_1-1_3-1_N_30} which attracts attention: While for small values of $c$ the tunneling period increases faster than exponential and shows an almost monotonic behavior, a sharp peak can be observed for $c\approx 1.98/(N+1)$ which in fact is a true divergence. This phenomenon is known as ``coherent destruction of tunneling'' (CDT) \cite{Gros91a, Gros91b, Grif98}. At these points in parameter space the two levels belonging to the tunneling doublet cross. Since they have different parity such crossings are not forbidden. This zero splitting automatically leads to a divergence in the tunneling period. Thus, the relatively high value of the tunneling period that can be observed in figure \ref{Figure:tunnelzeit_c_2_N_0_150} for $N=30$ particles is due to the direct neighborhood to the CDT divergence in parameter space. A diverging tunneling time means in fact total suppression of the dynamical many particle tunneling and therefore a real self-trapping of the condensate not only in the mean-field approximation, but also in the full many particle regime. \subsection{Tunneling landscapes} Let us finally explore the whole parameter space available for a symmetric two-mode system. One system parameter is the interatomic interaction strength $c$, the second parameter is the time scale of the system, which is given by the single particle tunneling coupling $v$. Both of the effects CDT and CAT are governed by the behavior of the quasi-energies in dependence on the system parameters, which in turn is ruled by the symmetries of the system. These are mappings under which the eigenvalues and symmetry properties of $F$ are preserved; for the Floquet operator \eqref{floq} they are given by orthogonal matrices. Therefore $F$ has codimension $n=2$, which means that at least two system parameters have to be varied to enforce a degeneracy of two quasi-energies belonging to the same symmetry class \cite{Wign59,Dyso62}. Hence accidental degeneracies of such quasi-energy planes $\epsilon_\kappa(c,v)$ occur at most at isolated points in the parameter space spanned by $c$ and $v$. However, planes belonging to \textit{different} symmetry classes, i.e.~having opposite parity, actually exhibit exact crossings which form a one dimensional manifold in the $(c,v)$-plane. Thus, the quasi-energy planes $\epsilon_\kappa(c,v)$ cross along lines \cite[sect. 6]{Grif98}. These lines play an important role in the context of tunneling since the tunneling period $T_{\rm tunnel}$ diverges there. Plotting $T_{\rm tunnel}$ as a function of $c$ and $v$ yields the ``tunneling landscape'' of the kicked double-well BEC. Figure \ref{Figure:landsc0} shows $T_{\rm tunnel}(c,v)$ in false color representation on a logarithmic scale. \begin{figure} \centering \includegraphics[scale=1]{eps/v_0_4-1_2-c-kreuzung-HDimXklein.eps} \caption{Tunneling landscape, $T_{\rm tunnel}$ over $(c,v)$ in false color representation on logarithmic scale; $N=30$ particles, $\varepsilon = 0$ and $\tau=1$. The sharp dark lines with reduced tunneling times represent CAT regions whereas these with high values of the tunneling time represent CDT regions.} \label{Figure:landsc0} \end{figure} One observes indeed a rich landscape with diversified sequences of ridges and valleys. The sharp lines with small values of the tunneling period (dark blue) correspond to the CAT regions; along these lines avoided crossings of the quasi-energy planes with significant splitting occur. On the other hand, the sharp lines with high values of $T_{\rm tunnel}$ are actually lines of divergence since the quasi-energy planes intersect along these lines in parameter space. Thus, they correspond to the CDT regions. In the upper left corner of the presented section of parameter space, i.e., for smaller values of $c$ and larger values of $v$, regular dynamics prevails and almost no crossings of the quasi-energy planes take place. The consequence of this is a flat tunneling landscape area. The opposite holds in the chaotic regime; here a larger number of CAT resonances and CDT divergences occur. Remarkable are the ``avoided crossings'' between both the ridges and the valleys of the landscape. One of these is shown in figure \ref{Figure:landsc1} which is a magnification of figure \ref{Figure:landsc0}. \begin{figure} \centering \includegraphics[scale=1]{eps/v-c-kreuzungA3-HDim.eps} \caption{Tunneling landscape, blow-up of figure \ref{Figure:landsc0}; avoided crossing of two CDT ridges, in the gap one observes a CAT valley.} \label{Figure:landsc1} \end{figure} A systematical feature at these crossings is the direct neighborhood of resonances and divergences which could already be observed in figure \ref{Figure:tunnel_c_1-1_3-1_N_30} that is actually a cut through figure \ref{Figure:landsc0} at $v=1$. Another interesting fact about this tunneling landscape is, that in principle the system can be tuned to the CDT regime by adjusting one parameter. Thus for this kicked system real self-trapping is possible not only in the mean-field regime, but also for the full quantum system and the system can be tuned from enforced tunneling to this regime by rather slight variations of the parameters. This yields the possibility of a systematic and accurate population transfer between the two potential wells. \section{Conclusions} In the present paper we investigated a two-mode BEC with a periodically kicked interaction term in a many particle as well as in a mean-field description. This system is actually equivalent to the kicked top which is a standard example of quantum chaos. Besides the new possible experimental realization of the kicked top, this new context sheds light on different aspects of the model. We showed that the tunneling oscillations of the many particle system which usually suppress the self-trapping effect show a rich spectrum of phenomena like coherent destruction of tunneling and chaos assisted tunneling. The sensitive parameter dependence of the tunneling behavior of the system offers an interesting tool for the manipulation of Bose-Einstein condensates. \section*{References} \begin{thebibliography}{10} \bibitem{Ande95} M.~H. Anderson, J.~R. Ensher, M.~R. Matthews, C.~E. Wieman, and W.~E. Cornell, Science {\bf 269} (1995) 198 \bibitem{Davi95} K.~B. Davis, M.~O. Mewes, M.~R. Andrews, N.~J. van Druten, D.~S. Durfee, D.~M. Kurn, and W.~Ketterle, Phys. Rev. Lett. {\bf 75} (1995) 3969 \bibitem{Albi05} M.~Albiez, R.~Gati, J.~F\"olling, S.~Hunsmann, M.~Cristiani, and M.~K. Oberthaler, Phys. Rev. Lett.{\bf 95} (2005) 010402 \bibitem{Bloc05} I.~Bloch, Nature Physics {\bf 1} (2005) 23 \bibitem{Milb97} G.~J. Milburn, J.~Corney, E.~M. Wright, and D.~F. Walls, Phys. Rev. A {\bf 55} (1997) 4318 \bibitem{Holt01a} M.~Holthaus and S.~Stenholm, Eur. Phys. J. B {\bf 20} (2001) 451 \bibitem{Holt01b} M.~Holthaus, Phys. Rev. A {\bf 64} (2001) 011601 \bibitem{Angl01} J.~R. Anglin and A.Vardi, Phys. Rev. A {\bf 64} (2001) 013605 \bibitem{Mahm05} K.~W. Mahmud, H.~Perry, and W.~P. Reinhardt, Phys. Rev. A {\bf 71} (2005) 023615 \bibitem{Moss06} S.~Mossmann and C.~Jung, Phys. Rev. A {\bf 74} (2006) 033601 \bibitem{Wu06} Biao Wu and Jie Liu, Phys. Rev. Lett. {\bf 96} (2006) 020405 \bibitem{06semiMP} H.~J. Korsch and E.~M. Graefe, Phys. Rev. A {\bf \phantom{0}} (2007) in press (preprint: quant--ph/0611040) \bibitem{Stoe99} H.-J. St\"ockmann,{\em Quantum Chaos}, Cambridge University Press, Cambridge, 1999 \bibitem{Haak01} F.~Haake, {\em Quantum Signatures of Chaos}, Springer, Berlin, Heidelberg, New York, 2001 \bibitem{Haak86} F.~Haake, M.~Ku\'s, and R.~Scharf, Z. Phys. B{\bf 65} (1986) 381 \bibitem{Haak00} F.~Haake, Journal of Modern Optics {\bf 47} (2000) 2883 \bibitem{Xie05} Q.~Xie and W.~Hai, Eur. Phys. J. D {\bf 33} (2005) 265 \bibitem{Zhan90} W.-M. Zhang, D.~H. Feng, and R.~Gilmore, Rev. Mod. Phys. {\bf 62} (1990) 867 \bibitem{Arec72} F.~T. Arecchi, E.~Courtens, R.~Gilmore, and H.~Thomas, Phys. Rev. A {\bf 6} (1972) 2211 \bibitem{Gilm75} R.~Gilmore, C.~M. Bowden, and L.~M. Narducci, Phys. Rev. A {\bf 12} (1975) 1019 \bibitem{Pere86} A.~M. Perelomov, {\em Generalized Coherent States and Their Applications}, Springer, Berlin Heidelberg New York London Paris Tokyo, 1986 \bibitem{07meanf} D.~Witthaut, F.~Trimborn, and H.~J. Korsch, {\bf \phantom{0}} (2007) (preprint: cond--mat/0701383) \bibitem{Grif98} M.~Grifoni and P.~H\"anggi, Phys. Rep.{\bf 304} (1998) 229 \bibitem{Bayf99} J.~E. Bayfield, {\em Quantum Evolution}, John Wiley and Sons, New York, 1999 \bibitem{Lin90} W.~A. Lin and L.~E. Ballentine, Phys. Rev. Lett. {\bf 65} (1990) 2927 \bibitem{Lin92} W.~A. Lin and L.~E. Ballentine, Phys. Rev. A {\bf 45} (1992) 3637 \bibitem{Bavl93} R.~Bavli and H.~Metiu, Phys. Rev. A {\bf 47} (1993) 3299 \bibitem{95tun} V.~Averbuckh, N.~Moiseyev, B.~Mirbach, and H.~J. Korsch, Z. Phys. D {\bf 35} (1995) 247 \bibitem{Bonc98} L.~{Bonci, A. Farusi, P. Grigolini and R. Roncaglia}, Phys. Rev. E {\bf 58} (1998) 5689 \bibitem{Aver02} V.~Averbukh, Shmuel Osovski, and N.~Moiseyev, Phys. Rev. Lett. {\bf 89} (2002) 253201 \bibitem{Hens00} W.~K. Hensinger, A.~G. Truscott, B.~Upcroft, N.~R. Heckenberg, and H.~Rubinsztein-Dunlop, J. Opt. B {\bf 2} (2000) 659 \bibitem{Hens01a} W.~K.~Hensinger et. al., Nature {\bf 412} (2001) 52 \bibitem{Hens01b} W.~K.~Hensinger et. al., Phys. Rev. A {\bf 64} (2001) 033407 \bibitem{Stec01} D.A. Steck, W.~H. Oskay, and M.~G. Raizen, Science {\bf 293} (2001) 274 \bibitem{Salm02} G.L. Salmond, C.~A. Holmes, and G.~J. Milburn, Phys. Rev. A {\bf 65} (2002) 033623 \bibitem{Osov05} S.~Osovski and N.~Moiseyev, Phys. Rev. A {\bf 72} (2005) 033603 \bibitem{Husi40} K.~Husimi, Proc. Phys. Math. Soc. Japan {\bf 22} (1940) 264 \bibitem{Hine05} A.~P. Hines, R.~H. McKenzie, and G.~J. Milburn, Phys. Rev. A {\bf 71} (2005) 042303 \bibitem{Shin94} J.~Y. Shin and H.~W. Lee, Phys. Rev. E {\bf 50} (1994) 902 \bibitem{Plat92} J.~Plata and J.~M.~Gomez Llorente, J. Phys. A {\bf 25} (1992) L303 \bibitem{Uter94} R.~Utermann, T.~Dittrich, and P.~H\"anggi, Phys. Rev. E{\bf 49} (1994) 273 \bibitem{Haen94} P.~H\"anggi, R.~Utermann, and T.~Dittrich, Physica B{\bf 194-196} (1994) 1013 \bibitem{brod01} O.~Brodier, P.~Schlagheck, and D.~Ullmo, Phys. Rev. Lett. {\bf 87} (2001) 064101 \bibitem{Elts05} C.~Eltschka and P.~Schlagheck, Phys. Rev. Lett. {\bf 94} (2005) 014101 \bibitem{Gros91a} F.~Grossmann, T.~Dittrich, P.~Jung, and P.~H\"anggi, Phys. Rev. Lett.{\bf 67} (1991) 516 \bibitem{Gros91b} F.~Grossmann, P.~Jung, T.~Dittrich, and P.~H\"anggi, Z. Phys. B{\bf 84} (1991) 315 \bibitem{Toms94} S.~Tomsovic and D.~Ullmo, Phys. Rev. E {\bf 50} (1994) 145 \bibitem{Kohl98} S.~{Kohler, R. Utermann, P. H\"anggi and T. Dittrich}, Phys. Rev. E {\bf 58} (1998) 7219 \bibitem{Wign59} E.~P. Wigner, {\em Group Theory and its Applications to the Quantum Mechanics of Atomic Spectra}, Academic, New York, 1959 \bibitem{Dyso62} F.~J. Dyson, J. Math. Phys. {\bf 3} (1962) 140, 157, 166 \end{thebibliography} %\bibliographystyle{unsrtot} %\bibliography{abbrev,dipldiss,paper60,paper70,paper80,paper90,paper00,publko,rest} \end{document} }
\caption{a) Mean-field dynamics for the parameters $\varepsilon = 0$, $v=1$, $\tau = 1$ and $c=2/(N+1)$, the fixed points $\boldsymbol s_\pm$ are marked with red dots. b) Propagation of many-particle state $\left|-\right>$ located on the southern fixed point for $N=10$ particles; the dynamical tunneling to the northern fixed point lasts $40$ periods here; c) as before, but $N = 20$ particles; tunneling during $1000$ periods d) as in c), but $N=100$ particles, dynamical tunneling cannot be observed here.} \label{Figure:tunnel_c_2_fix} \end{figure} It should be emphasized, that here the dynamical tunneling between both of the island states is a coherent many particle process and may not be confused with the single particle tunneling which causes the Rabi oscillations. Nevertheless it corresponds to an actual tunneling of the condensate between the potential wells, since the northern state is localized almost in the first well, whereas the southern state occupies mostly the second well. In figure \ref{Figure:tunnel_c_2_fix} the tunneling process starting from the fixed point on the southern hemisphere in the state $\left|\psi(0)\right>= \left|-\right>$ is shown for several parameter sets. An interaction strength of $c = 2/(N+1)$ is chosen to ensure the two regular islands to be not too small and well separated. One observes dynamical tunneling through the Bloch sphere. The normalized expectation value $\left<\boldsymbol L\right>/\ell$ spirals its way on a tube encircling the straight line connecting the two island states to the other side of the sphere. The radius of the tunneling tube is determined by the uncertainty of the direction of the angular momentum vector in a coherent state, which is proportional to $1/N$. The more particles one puts into the system, the tighter is the spiral around the connection line of the fixed points and the longer gets the tunneling time. However, under some conditions the tunneling time can also decrease with a growing number of particles, leading to a strongly enhanced population transfer, as will be explained in section \ref{subsecparameter}. But let us first have a closer look at the tunneling process. The squared absolute values $\left|\left<\psi(t)\big|\pm\right>\right|^2$ of the projections of $\left|\psi(t)\right>$ on the coherent island states and the orthogonal subspace for $\left|\psi(0)\right>=\left|-\right>$ for case c) of figure \ref{Figure:tunnel_c_2_fix} with $N = 20$ is shown in figure \ref{Figure:inselproj_N_20_c_2} as a function of time. \begin{figure} \centering \includegraphics[scale=1]{eps/inselproj_N_20_c_2.eps} \caption{Squared absolute values of the projections of $\left|\psi(t)\right>$ on the coherent states $\left|\pm\right>$ during propagation over $1000$ periods for $N=20$ particles, $\varepsilon = 0$, $v=1$, $\tau=1$ and $c=2/(N+1)$; $\textcolor{blue}{-}$ $\left|\left<-\big|\psi(t)\right>\right|^2$, $\textcolor{green}{-}$ $\left|\left<+\big|\psi(t)\right>\right|^2$ and $\textcolor{red}{-}$ orthogonal subspace.} \label{Figure:inselproj_N_20_c_2} \end{figure} After $1000$ periods, the population has been transferred almost completely to the state $\left|+\right>$ on the northern hemisphere. For all particle numbers $N$ considered in figure \ref{Figure:tunnel_c_2_fix}, the population $\left|\left<\kappa\big|-\right>\right|^2$ of the southern coherent island state on the Floquet states $\left|\kappa\right>$ is significantly large only for two of the $(N+1)$ Floquet states which are denoted by $\left|\kappa_\pm\right>$ with corresponding quasi-energies $\epsilon_\pm$. As indicated by the $\pm$-sign, they belong to different symmetry classes and the island states can be reconstructed up to 94\,\% by the linear combinations \begin{equation}\label{k4_linkomb} \left|\psi_\pm\right> = \frac{1}{\sqrt{2}}\left(\left|\kappa_+\right> \mp \left|\kappa_-\right>\right) \end{equation} which hence are a good approximation for the coherent states $\left|\pm\right>\approx \left|\psi_\pm\right>$. If one chooses $\left|\psi_-\right>$ as a starting vector, $\left|\psi(t)\right>$ always stays a linear combination of the $\left|\kappa_\pm\right>$ during propagation. Remembering equation \eqref{k3_Feigenzu} for the decomposition of the Floquet states, the propagation over $n$ kick periods yields \begin{eqnarray} \left|\psi(n\tau)\right> &=& \frac{1}{\sqrt{2}}{\rm e}^{-{\rm i}n\epsilon_+\tau }\left(\left|\kappa_+\right> + {\rm e}^{-{\rm i}n(\epsilon_- - \epsilon_+)\tau}\left|\kappa_-\right>\right)\\ &=& \frac{1}{2}{\rm e}^{-{\rm i}n\epsilon_+\tau }\left( \left(1 - {\rm e}^{-{\rm i}\phi_n} \right)\left|+\right> + \left(1 + {\rm e}^{-{\rm i}\phi_n} \right)\left|-\right> \right), \end{eqnarray} with $\phi_n = n(\epsilon_- - \epsilon_+)\tau$. The expectation value of the angular momentum varies as \begin{equation}\label{ellipse} \left<\boldsymbol L (n\tau) \right> = \frac{1-\cos{\phi_n}}{2}\boldsymbol L_+ + \frac{1+\cos{\phi_n}}{2}\boldsymbol L_- + \sin{\phi_n}{\rm Im}{\boldsymbol L_{-+}} \end{equation} with $ \boldsymbol L_\pm = \left<\pm\right|\boldsymbol L \left|\pm\right>$, $ \boldsymbol L_{-+} = \left<-\right|\boldsymbol L \left|+\right>$, where in the present cases ${\rm Im}\boldsymbol L_{-+}$ is zero or very small so that the ellipse \eqref{ellipse} degenerates to a straight line connecting the island centers at $\boldsymbol L_-$ and $\boldsymbol L_+$. Thus the spiraling on a tube around this line is due to the other contributing Floquet states. We conclude, that the subsystem consisting of the two island states $\left|\pm\right>$ can be described in a good approximation by a two state system of the Floquet states $\left|\kappa_\pm\right>$. Thus, as any two level system, the system exhibits periodic tunneling with period \begin{equation}\label{k4_tunnelzeit} T_{\rm tunnel} = \frac{2\pi}{\epsilon_- - \epsilon_+} = \frac{2\pi}{\Delta\epsilon}. \end{equation} These considerations are entirely equivalent to the tunneling process through a symmetric potential barrier, where the quasi-energies adopt the function of the energies of time independent systems. Thus the tunneling period is solely ruled by the quasi-energy splitting $\Delta\epsilon$. One should keep in mind, that despite the similarities in the formalism this is no single particle, but a coherent many particle process. Still the dynamical tunneling between the two island states corresponds to an actual tunneling of the BEC from one potential well to the other. If the system is localized in the southern island $\left|-\right>$, the expectation value of $L_z$ is negative and the condensate is located mainly in the second well; in the case $c = 2/(N+1)$ over $80\,\%$ of the condensate is trapped there. The opposite is true for the northern island $\left|+\right>$. To give an impression of the localization of the Floquet states, one can compute quantum phase space distributions, e.g.~the Husimi distributions \cite{Husi40, Hine05} \begin{equation} Q(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi\right>\right|^2. \end{equation} via the $SU(2)$ coherent states $\left|\vartheta,\varphi\right>$. In figure \ref{Figure:husimi-insel}, the Husimi distribution for the linear combinations $\left|\psi_\pm\right>$ of the two Floquet tunneling states are shown. \begin{figure} \centering \includegraphics[scale=1]{./eps/hus_minus_c_2_N_30.eps}\includegraphics[scale=1]{./eps/hus_plus_c_2_N_30.eps} \caption{Husimi distributions of the island states $\left|\psi_\pm\right>$; left: $Q_-(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi_-\right>\right|^2$, right: $Q_+(\vartheta,\varphi) = \left|\left<\vartheta,\varphi\big|\psi_+\right>\right|^2$ for $N=20$ particles, $\varepsilon = 0$, $v=1$, $\tau=1$ and $c=2/(N+1)$.}
\caption{Probability distribution of the potentials of nodes with degree $k$ for SF network with $m = 3$ before (inset) and after optimization. The degrees are binned to indicate the correlation between degree of a node and its potential: $3 < k \leq 10$ ({\color{red}$+$}), $10 < k\leq 30$ ($\square$), $30 < k \leq 50$ ({\color{blue} $\bullet$}), $50 < k\leq 100$ ({\color{magenta} $\circ$}). Before the optimization $P(V_i)$ do not show any correlations with $k$, however after the optimization, nodes with high degree, get the large values of potential which facilitates reduced jamming. \label{fig:fig4}}
\caption{{\it Left:} \gray{} flux vs.\the DM central spike power-law index. The lines are shown for a series of annihilation cross sections$\langle\sigma v\rangle$. {\it Right:} The visual K-band magnitude of DM burners at the GC without extinction vs.\the effective surface temperature.}
\caption{{\it Left:} The ratio of anisotropic IC to isotropic IC for Galactic longitudes $l=0^\circ$ and 180$^\circ$ vs.\Galactic latitude.{\it Right:} \gray{} spectrum of inner Galaxy ($330^\circ<l<30^\circ, |b|<5^\circ$) for an optimized model. Vertical bars: COMPTEL and EGRET data, heavy solid line: total calculated flux. This is an update of the spectrum shown in \cite{2004ApJ...613..962S}.}
\caption{(a) Time evolution for the AB model of the average interface density on different realizations of the network with 2500 agents; 20000 runs on each (empty symbols). The extreme cases were selected as examples of networks where {\it trapped metastable states} (see text) are found {\it often} (\textcolor{red}{$\bigcirc$}); and found {\it rarely} (\textcolor{red}{$\triangle$}). For comparison, the average over 500 networks (10 runs on each) is also shown ($\blacksquare$). Inset: time evolution for the voter model of the average interface density for four realizations of the networks of 2500 agents; 5000 runs on each network. (b) Time evolution of the interface density in single realizations of the AB dynamics on a network with 2500 agents. A class of realizations decay to the absorbing state after a coarsening stage (solid black lines), while others fall in long lived trapped metastable states. The latter display several plateaus, indicating hierarchical levels of ordering before reaching the absorbing state, or cascading between several trapped metastable states. }
\caption{Lattice parameters of PZT 20/80: bulk ($\circ$) and thin film (out- \textcolor{blue}{$\blacktriangle$} and in-plane, a \textcolor{red}{$\blacktriangledown$} and b \textcolor{orange}{$\blacktriangledown$}) deposited on SrTiO$_3$ ($\vardiamondsuit$)}
\caption{The spectrum of \grays\from outer (monoenergetic protons of 25 TeV, blue-dots) and inner clouds (power-law with index$-2.29$, red-dashes); normalisations are arbitrary. Solid line shows the total spectrum. Data: HESS \cite{Aharonian2006a}. \label{cap:gamma-ray-flux}}
\caption{Beam/target/detector setup for simulating CR interactions in moon rock. The primary beam enters the moon rock target with incident polar angle $\theta_p$. Secondary \gray{s} are emitted with polar angle $\theta$. The detection volume surrounds the target.}
\caption{\gray{} yield per proton interaction integrated over all emission angles from the Moon surface. % for selected proton energies and %incident angles. Line-styles: red-dashed, $\cos\theta_p = 1$; blue-solid, $\cos\theta_p = 0.1$. Line-sets: lower, $E_p = 500$ MeV; upper, $E_p = 10000$ MeV.}
\caption{Calculated \gray{} albedo spectrum of the Moon. Line-styles: black-solid, total; blue-dotted, limb -- outer $5'$; red-dashed, centre -- inner $20'$. Upper solid line: $\Phi_0 = 500$ MV; lower solid line: $\Phi_0 = 1500$ MV. Data points from the EGRET \cite{Thompson1997} with upper and lower symbols corresponding to periods of lower and higher solar activity, respectively. The differential 1 year sensitivity of the LAT is shown as the hatched region.}
\caption{\label{phasefig} Steady-state distributions of the number of particles on a site from simulations of the non-conserving ZRP model on a fully connected lattice (open shapes) and a 1d lattice (character symbols), compared with theoretical curves (dashed lines). Simulations were run on a system with L=5000 lattice sites and $b=2.6$, $k=3$. Data are shown for: Critical Phase A, $s=2$ (${\color{red}\circ}$, ${\color{red}+}$ and {\color{red}- -}); Critical Phase B, $s=1.7$ (${\color{green}\Box}$, ${\color{green}\times}$ and {\color{green}- -}); Weak High Density Phase, $s=1.2$ (${\color{blue}\Diamond}$, ${\color{blue}*}$ and {\color{blue}- -}); Strong High Density Phase, $s=0.4$ (${\color{magenta}\triangle}$, ${\color{magenta}\bullet}$ and {\color{magenta}- -}).}
\caption{(color online) Dependence of the ground-state energy of \elem{O}{16} (a) and \elem{Ca}{40} (b) on the truncation parameter $\kappa_{\min}$ obtained with $\VO_{\UCOM}$. The three data sets correspond to model spaces with up to 2p2h (\symbolcircle[FGBlue]), 3p3h (\symboldiamond[FGRed]), and 4p4h states (\symbolbox[FGGreen]), respectively. The lines show a 5th order polynomial interpolation.}
\caption{(color online) Convergence of the ground-state energy for \elem{He}{4} (a) and \elem{O}{16} (b) versus $N_{\max}$ obtained using $\VO_{\UCOM}$. Shown are three data sets corresponding to model spaces with up to 2p2h (\symbolcircle[FGBlue]), 3p3h (\symboldiamond[FGRed]), and 4p4h-states (\symbolbox[FGGreen]), respectively. Black crosses ($+$) indicate the results of full NCSM calculations. Lines to guide the eye.}
\caption{Extragalactic \gray{} and irreducible background in the GLAST-LAT based on our Monte Carlo study. Data points: cyan-circle from \cite{Sreekumar1998}; black-triangle from \cite{Strong2004a}. Blue points with error bars: irreducible intensity from our analysis. Black-lines show the uncertainty on the irreducible intensity due to uncertainties in the charged particle flux model. Black-hatched region shows the combined uncertainty from the charged particle flux model and hadronic physics modelling. Red-shaded region shows the hatched region re-scaled to the nominal CGRO orbital altitude of 450 km.}
\caption{{\it Left:} The ratio of anisotropic IC to isotropic IC for Galactic longitudes $l=0^\circ$ and 180$^\circ$ vs.\Galactic latitude\cite{Moskalenko2007}. {\it Right:} \gray{} spectrum of inner Galaxy ($330^\circ<l<30^\circ, |b|<5^\circ$) for the optimised model. Vertical bars: COMPTEL and EGRET data, heavy solid line: total calculated flux. This is an update of the spectrum shown in \cite{Strong2004b}.}
\caption{\label{fig:2} Upper panel: measured Wannier-Stark tunnelling rates (\fullcircle) for $g \approx 0$ in a lattice of depth $V_0 \approx 3.5$. The full line shows the theoretical expectation obtained by the methods described in \cite{GKK,WSM}, while the dashed curve represents the Landau-Zener prediction $\Gamma_{\rm LZ}$. Lower panel: the ratio of the theoretical (\full) and the measured (\fullcircle) rates and $\Gamma_{\rm LZ}$. Even if the modulation here is relatively small due to the chosen small lattice depth, we note the extention of the measured rates over two orders of magnitude in the upper panel. The nicely resolved central peak around $F_0\approx 1.35$ corresponds to RET between second-nearest potential wells, while the left one at smaller $F_0\approx 0.5$ is a consequence of RET between third-nearest well. The shoulder on the right shows signatures of the next-nearest neighbour RET peak around $F_0 \approx 2.7$, corresponding to the situation sketched in figure~\ref{fig:1}. }
\caption{\label{fig:3} ($a$) same theoretical data as in the upper graph of figure~\ref{fig:2} with measured tunnelling rates for a small nonlinearity $C \approx 0.025$ (\fullsquare). The mean-field, repulsive nonlinearity washes out the RET peak structures of figure~\ref{fig:2}, and this can be used as an additional experimental handle to globally enhance the rates. ($b$) and ($c$) show numerical data from a Bose-Hubbard model of 7 atoms in 6 potential wells for $V_0=3$ (fixing the hopping constant $J\approx 0.22$), interaction constant $U=0.2$, and $F_0=0.47$ ($b$) or $F_0=0.16$ ($c$). The broad distribution at the larger force arises from the regular Bloch oscillation dynamics of the atoms. The distribution in ($c$) shows that there is a small number of preferred channels to tunnel to the first excited levels originating from the quantum chaotic motion of the atoms along the wells (leading e.g. to interaction-induced decoherence of the Bloch oscillations \cite{BK2003}). }
\caption{Integrated-flux LCs of \mbox{Mrk 501} for the flare nights of June 30 and July 9. Horizontal bars represent the 2-minute time bins, and vertical bars denote 1$\sigma$ statistical uncertainties. For comparison, the Crab emission is also shown as a lilac dashed horizontal line. The vertical dot-dashed line divides the data into 'stable' (i.e., pre-burst) and 'variable' (i.e., in-burst) emission emission. The horizontal black dashed line represents the average of the 'stable' emission. The solid black curve represents the best-fit flare model (see eq. \ref{eq_fitflare}). The insets report the fit parameters and goodness of the fit. The bottom plots show the mean background rate during each of the 2-minute bins of the LCs. The insets report the mean background rate during the entire night, resulting from a constant fit to the data points. The goodness of such fit is also given. The background rates are constant along the entire night. Consequently, the variations seen in the upper panels of the middle and right-hand plot correspond to actual variations of the \vhe\\gray\flux from Mrk501, thus ruling out detector instabilities and/or atmospheric changes.}
\caption{$v_2$ as a function of $N_{\rm part}$ at mid-rapidity for Pb-Pb collisions at LHC ($\sqrt{s_{NN}}=5.5$~TeV). \full and \dashddot: $\varepsilon$ scaling ($K=0$ in (\ref{v2k})); \dashed and \dotted: incl.\incomplete thermalization, with two values of the partonic cross section.\fullsquare: PHOBOS data for Au-Au collisions at RHIC~\cite{Back:2004mh}. The vertical scale is arbitrary (see text). \label{fig:v2lhc}}
\caption{\label{fig1}Storage capacity $\alpha$ as a function of the intensity of the external noise $D$ for various values of $\tilde{D}$. The solid curve denotes the storage capacity for $\tilde{D}=0$. The broken (\broken), dashed (\dashed) and dotted (\dotted) curves represent the storage capacity for $\tilde{D}=0.2$, $0.4$, $0.6$, respectively. The retrieval state locates below the curve and vanishes at $\tilde{D}\sim 1.05$. We set $A=20$. }
\caption{\label{Fig:Int_vs_NoInt}(a) Typical Rydberg atom numbers as a function of excitation time $\tau$ for $n_{\text{g}}=5.2\times 10^{19}\text{ m}^{-3}$, i.e., $N_{\text{g}}=1.1\times 10^7$ and $\Omega_0 = 2\pi\times$90.5 kHz. The error bar shows the statistical fluctuation of the Rydberg atom number over ten experiments at every excitation time. (b) Rydberg atom number $N_{\text{R}}$ (\textcolor{blue}{$\square$}) as a function of the number of ground state atoms $N_{\text{g}}$ and their density $n_{\text{g}}$ (upper scale). From a power-law fit $N_{\text{R}} \propto N_{\text{g}}^{0.43\pm 0.03}$ is deduced. For comparison the upper solid line shows the expected Rydberg number for an excitation of non-interacting Rydberg atoms. The lower solid line takes a frequency uncertainty of the excitation laser of 2$\pi\times$1.5 MHz into account. The filled data points (\textcolor{blue}{$\blacksquare$}) indicate the atom numbers for which the rotary echo signals are shown in \fig\ref{Fig:EchoExp}. The statistical fluctuation is smaller than the marker size.}
\caption{\label{Fig:VisAll} Visibility as a function of the peak density of ground state atoms $n_{\text{g}}$ for three different excitation times. Namely, $\tau=478$ ns (\textcolor{red}{$\square$}), $\tau=534$ ns (\textcolor{blue}{$\bigcirc$}) and $\tau=659$ ns ($\triangle$). All three data sets were fitted with a simple exponential decay to guide the eye. The error bars are obtained from the parabolic fits to the data (\fig{\ref{Fig:EchoExp}}).}
\caption{% The mass spectrum for the superpartners of the Standard Model for four different models (in four different colors): \SUSYLR$+$AMSB, \textcolor{red}{mAMSB}, \textcolor{green}{mSUGRA}, and \textcolor{blue}{mGMSB}. Note that for the \SUSYLR$+$AMSB, $\tilde{t}_2$ and $\tilde{b}_2$ are mostly right-handed; in contrast with the usual mSUGRA or mGMSB cases where they are typically mostly left-handed.% }
\caption[Indirect route: Results.]{(Color online). Results for the optimization of the indirect trajectory [see \fref{fig:pot2Dindirect}]. (a) Optimized laser field. (b) Spectrum of optimized laser field. Lower panel: $x$ component. Upper panel: $y$ component. (c) Time evolution of the most important occupation numbers: $|\langle n | \Psi(t) \rangle |^2 $: Top panel: $n=0$ (\full), $n= 4$ (\broken), $n=7$ (\chain). Bottom panel: $n=5$ (\dotted), $n=8$ (\dashddot), $n=9$ (\opensquare). The (\full) line corresponds to the occupation for all states with $n>15$. (d) Lower panel: Target trajectory [(\full) line], the expectation value of the position operator [(\dashed) line], and the position of the density maximum during the propagation (\opencircle). Upper panel: Convergence of the algorithm where the (\dashed) line corresponds to $J$, the (\full) line to $J_1$, and the laser fluence $E_0$ is depicted by the (\dotted) line.}
\caption[Direct route: Results.]{(Color online). Results for the optimization of the direct trajectory [see \fref{fig:pot2Ddirect}]. (a) Optimized laser field. (b) Spectrum of optimized pulse. The $x$ component [(\full) line] and the $y$ component [(\dashed) line] lie on top of each other. (c) Time evolution of the most important occupation numbers: $|\langle n | \Psi(t) \rangle |^2 $: top panel: $n=0$ (\full), $n= 3$ (\broken), $n=9$ (\opensquare), bottom panel: $n=5$ (\dotted), $n=6$ (\chain), $n=10$ (\opencircle). The (\full) line corresponds to the occupation for all states with $n>15$. (d) Lower panel: (\dashed) Expectation value of the position operator which lies on top of the target trajectory [(\full) line], (\opencircle) position of the density maximum during the propagation. Upper panel: Convergence of the algorithm. The (\dashed) line corresponds to $J$, the (\full) line to $J_1$, and the (\dotted) line to the laser fluence $E_0$.}
\caption{Spin wave dispersion of magnetite above $T_{V}.$ \Black (gray) symbols are inelastic neutron scattering data from MAPS (HB-3).\Lines are results from a Heisenberg model with parameters discussed in the main text. The hatched area contains very broad B-site spin waves.}
\caption{Inelastic neutron scattering intensity as a function of energy transfer for cuts along;\(a) $[H20]$, (b) $[H30]$, and (c) $[H40]$.\Blue (red)\symbols indicate measurements done at$T=$ 110 K (130 K). Blue (red) lines are Gaussian fits to the 110 K (130 K) data. The dashed line is the background estimated from cuts along $[H60]$. Panels (d) - (f) show the identical cuts as (a) - (c) as calculated from a Heisenberg model for magnetite with $J_{BB}=0.44$ meV (blue) and $J_{BB}=0.69$ meV (red).\Spin wave modes are labelled by symmetry.}
\caption{(a) Neutron structure factor for B-site optical spin waves summed from 75-90 meV. \Red (blue) circles are data at 130 K (110 K).\Red (blue) lines are calculated from a Heisenberg model with$J_{BB}=$0.69 (0.44)\meV.\Horizontal bars indicate dominant symmetry.\(b) Structure factor for A-site optical spin waves from 115 -120 meV. \Black line is the result of Heisenberg model.}
\caption{\label{transposeYD} (a) Transposed Young diagram for the unrestricted partition $91=18+16+13+13+9+6+5+5+3+3$. (b) Represents (a) in terms of non-interacting bosons (represented by \opencircle) occupying energy levels $\epsilon_i=i$ for $i=1,2,\ldots,10$. (c) The configuration obtained from the bosonic configuration (b) by transferring particles from the higher levels to the lower levels such that in the final configuration all the levels below the highest occupied level $\epsilon_5=5$ receive 2 new particles (represented by \fullcircle) each, where \opencircle represents the particles originally present in the initial bosonic configuration. (d) The Young diagram corresponding to the configuration (c). This is the transposed Young diagram of the partition $49=18+14+9+7+1$, in the minimal difference 2 partition problem.}
\caption{Spectrum of the full operator and reduced-order models for rank (a) 4, (b) 8, (c) 30. The BPOD modes are from the eight-mode output projection. Symbols: \textcolor{red}{BPOD ($\square$)}, \textcolor{blue}{POD ($\bigcirc$)}, full operator ($+$) Only the most important part of the full spectrum is shown.}
\caption{(a) $\alpha=1,\beta=1$ optimal perturbation at $Re=1000$, eight-mode output projection, 4-mode and 8-mode models. Full simulation ($+$), \textcolor{blue}{4-mode POD ($\bigcirc$)}, \textcolor{red}{4-mode BPOD ($\square$)}, \textcolor{green}{8-mode POD ($\diamond$)}, \textcolor{magenta}{8-mode BPOD ($+$)} (b) First two outputs, symbols as defined in (a).}
\caption{Left: The comparison of spectra of the full operator at $\alpha=1,\beta=1$ to the spectra of rank 12 models as the value of $Re$ is increased to 2000. Right: The performance of corresponding reduced-order models at $Re=2000$ - first two outputs. See text for detailed description. Symbols: full ($\times$), \textcolor{red}{BPOD ($\square$)}, \textcolor{blue}{POD ($\bigcirc$)}.}
\caption{\label{fig:LoRnvsL} Phase diagram of all superconducting (\textcolor{blue}{$\circ$}) and insulating (\textcolor{red}{{\tiny $\blacksquare$}}) wires in $L/R_N$-$L$ space. Dashed line is $R_N = R_Q$.}
\caption{\label{fig:RnoLvsRn} Phase diagram of all superconducting (\textcolor{blue}{$\circ$}) and insulating (\textcolor{red}{{\tiny $\blacksquare$}}) wires in $R_N/L$-$R_N$ space with the separatrix (dashed line) given in the text.}
\caption{(Color online) The superconducting transition temperature ($T_c$) as measured by magnetometry as a function of the $\alpha$-decays per atom for this study (samples {\bf A} and {\bf B}) and in the literature (1 yr=4.9$\times10^{-6}$~$\alpha$-decays per atom for these samples). A linear decay of $T_c$ (\textcolor{red}{---}) starting from about 19.0 K and decaying at a rate of about 5.4$\times10^{5}$ K per $\alpha$-decay per atom is also shown for reference. The inset shows the crystal structure of PuCoGa$_5$.\cite{Sarrao02} }
\caption{(Color online) EXAFS data $k^3\chi(k)$ vs. $k$ from the (a) Pu $L_\textrm{III}$ edge, (b) the Co $K$ edge, and (c) the Ga $K$ edge, for a fresh ({\bf ---}, $0.2\times10^{-6}$~$\alpha$-decays per atom) and an aged ({\bf \textcolor{green}{---}}, 14.5$\times10^{-6}$~$\alpha$-decays per atom) sample.}
\caption{(Color online) Fourier transform (FT) of the $k^3\chi(k)$ data of the fresh sample ({\bf ---}) in Fig. \ref{ks_fig}, together with a fit to these data (\textcolor{red}{{\bf -~-~-}}) . The fit quality is such that the fit is difficult to distinguish from the data. The outer envelope is $\pm$ the transform amplitude and the inner modulating line is the real part of the complex transform. The Pu edge data (a) are transformed between 2.5-16.0 \AA$^{-1}$, Gaussian broadened by 0.3 \AA$^{-1}$, and are fit between 2.0 and 5.0 \AA{}. The Co edge data (a) are transformed between 2.5-11.0 \AA$^{-1}$, Gaussian broadened by 0.3 \AA$^{-1}$, and are fit between 2.0 and 5.5 \AA{}. The Ga edge data (a) are transformed between 2.5-15.0 \AA$^{-1}$, Gaussian broadened by 0.3 \AA$^{-1}$, and are fit between 1.6 and 5.0 \AA{}. }
\caption{(Color online) FT of the $k^3\chi(k)$ EXAFS data from the (a) Pu $L_\textrm{III}$, (b) Co $K$, and the (c) Ga $K$ edges are shown for sample ages about one year apart. Samples have accumulated doses of 0.2$\times 10^{-6}$~$\alpha$-decays per atom (---), 5.4$\times 10^{-6}$~$\alpha$-decays per atom (\textcolor{red}{---}), 10.2$\times 10^{-6}$~$\alpha$-decays per atom (\textcolor{blue}{---}), and 14.5$\times 10^{-6}$~$\alpha$-decays per atom (\textcolor{green}{---}). Transform ranges are between 2.5-16.0 \AA$^{-1}$, 2.5-10.0 \AA$^{-1}$, and 2.5-14.5 \AA$^{-1}$, respectively, all Gaussian broadened by 0.3 \AA$^{-1}$.}
\caption{The Hankel singular values $\sigma_j$: exact balanced truncation(\textcolor{green}{$\square$}); balanced truncation by the method of snapshots but without output projection(\textcolor{magenta}{$\circ$}); balanced POD with $r_{op}=1$ (\textcolor{cyan}{$\lozenge$}); balanced POD with $r_{op}=2$ (\textcolor{red}{$*$}); balanced POD with $r_{op}=6$ (\textcolor{blue}{$+$}); balanced POD with $r_{op}=10$ (\textcolor{black}{$\times$}). }
\caption{Error $||\tilde{G}-\tilde{G}_{\tilde{r}}||_\infty/||\tilde{G}||_\infty$, for exact balanced truncation(\textcolor{green}{$\square$}), balanced truncation by the method of snapshots but without output projection(\textcolor{magenta}{$\circ$}), balanced POD with $r_{op}=1$ (\textcolor{cyan}{$\lozenge$}), balanced POD with $r_{op}=2$ (\textcolor{red}{$*$}), balanced POD with $r_{op}=6$ (\textcolor{blue}{$+$}), and the lower bound for any model reduction scheme (\textcolor{black}{$-$}).}
\caption{Time varying $T$-periodic output projections versus time-invariant output projections: Error $||\tilde{G}-\tilde{G}_{\tilde{r}}||_\infty/||\tilde{G}||_\infty$, for balanced POD with $r_{op}=1$ (\textcolor{cyan}{$\lozenge$}), balanced POD with $r_{op}=2$ (\textcolor{red}{$*$}) and balanced POD with $r_{op}=6$ (\textcolor{blue}{$+$}). Solid lines correspond to cases using $T$ different projection matrices along one period for the periodic system, and dashed lines using one single projection matrix. The black solid line is the lower bound for any model reduction scheme (\textcolor{black}{$-$}).}
\caption{Mean square radius of gyration $\langle R_G^2 \rangle_L$ and $\langle R_G^2 \rangle_R$ for dynamics A ( no bond crossing, nbc) and B (allowed bond crossings, bc). For dynamics A, data points of linear polymers are shown by \opentriangle, where %$A_L=0.072 \pm 0.0005$ and $\nu_L= 1.2879 \pm 0.002$; %$A_L=0.072 \pm 0.001$ and $\nu_L= 1.288 \pm 0.002$; ring polymers by \fullsquare, where $A_R=0.043 \pm 0.0001$ and $\nu_R= 1.270 \pm 0.001$. %$A_R=0.043 \pm 0.0001$ and $\nu_R= 1.2696 \pm 0.001$. For dynamics B, linear polymers by \fulldiamond, where $A_L=0.175 \pm 0.001$ and $\nu_L= 1.204 \pm 0.002$; %$A_L=0.1751 \pm 0.0010$ and $\nu_L= 1.2041 \pm 0.002$; ring polymers by \opentriangledown, where $A_R=0.105 \pm 0.001$ and $\nu_R= 1.179 \pm 0.001$. %$A_R=0.1052 \pm 0.0004$ and $\nu_R= 1.1785 \pm 0.001$. }
\caption{The ratio $g= \langle R_G^2 \rangle_R / \langle R_G^2 \rangle_L$ versus $N$ with the fitting curve (\ref{eq:g-fit}). For dynamics A shown by \fulldiamond (nbc), we have $g_{\infty} = 0.559 \pm 0.007$, $B_g=0.402 \pm 0.438$ and $\Delta_g=1.173 \pm 0.706$. %$g_{\infty} = 0.5588 \pm 0.0068$, %$B_g=0.4019 \pm 0.4378$ and $\Delta_g=1.1728 \pm 0.7060$. Here $\chi^2=3.3$ for 8 data points. For dynamics B shown by \fulltriangle (bc), we have $g_{\infty} = 0.535 \pm 0.002$ $B_g=0.204 \pm 0.079$ and $\Delta_g=0.565 \pm 0.476$. %$g_{\infty} = 0.5352 \pm 0.0020$ %$B_g=0.2041 \pm 0.0788$ and $\Delta_g=0.5649 \pm 0.4755$. Here $\chi^2=14.9$ for 8 data points. }
\caption{Diffusion constants of ring and linear polymers for dynamics A (nbc) depicted by \opentriangle and \fullsquare, respectively, for dynamics B (bc) \fulldiamond and \opentriangledown, respectively. The horizontal axis denotes the number of segments, $N$. }
\caption{ Ratio $C=D_R/D_L$ versus $N$. For dynamics A (\fulldiamond) with fitting curve (\ref{eq:C-fit}), $C_{\infty} = 1.14 \pm 0.01$, $B_C= 0.094 \pm 76.4$ and $\Delta_C =2.98 \pm 454.97$. Here $\chi^2=6.0$ for 8 data points. For dynamics B (\opentriangledown), $C_{\infty} = 1.11 \pm 0.01$, $B_C= 161.8 \pm 2010.3$ and $\Delta_C =4.78 \pm 6.94$. Here $\chi^2=5$ for 8 data points.}
\caption{Beam/target/detector setup for simulating CR interactions in moon rock. The primary beam enters the moon rock target with incident polar angle $\theta_p$. Secondary \gray{s} are emitted with polar angle $\theta$. The detection volume surrounds the target.}
\caption{\gray{} yield per proton interaction integrated over all emission angles from the Moon surface. Left panel: yield calculated for $E_p = 10000$ MeV and $\cos\theta_p = 0.1$ with components shown. Line styles: blue-dashed, decay; red-long-dash-dot, bremsstrahlung; magenta-short-dash-dot, non-elastic scattering and de-excitation; cyan-dot, low energy neutron capture; black-solid, total. For each component, the initial process responsible for the production of the \gray\secondary is given by the line style but the yield distribution includes also processes such as Compton scattering. Right panel: yield calculated for two different energies and incident angles. Line styles: blue-solid,$\cos\theta_p = 0.1$; red-dashed, $\cos\theta_p = 1$. Line-sets: lower, $E_p = 500$ MeV; upper, $E_p = 10000$ MeV.}
\caption{Calculated \gray{} albedo spectrum of the Moon. Line-styles: black-solid, total; blue-dotted, limb -- outer $5'$; red-dashed, center -- inner $20'$. Upper solid line: $\Phi = 500$ MV; lower solid line: $\Phi = 1500$ MV. Limb and center components are only shown for $\Phi = 1500$ MV. Data points from the EGRET \citep{Thompson1997} with upper and lower symbols corresponding to periods of lower and higher solar activity, respectively. The differential 1 year sensitivity of the LAT is shown as the shaded region.}
\caption{Beam/target/detector setup for simulating CR interactions in moon rock. The primary beam enters the moon rock target with incident polar angle $\theta_p$. Secondary \gray{s} are emitted with polar angle $\theta$. The detection volume surrounds the target.}
\caption{\gray{} yield per proton interaction integrated over all emission angles from the Moon surface. Left panel: yield calculated for $E_p = 10000$ MeV and $\cos\theta_p = 0.1$ with components shown. Line styles: blue-dashed, decay; red-long-dash-dot, bremsstrahlung; magenta-short-dash-dot, non-elastic scattering and de-excitation; cyan-dot, low energy neutron capture; black-solid, total. For each component, the initial process responsible for the production of the \gray\secondary is given by the line style but the yield distribution includes also processes such as Compton scattering. Right panel: yield calculated for two different energies and incident angles. Line styles: blue-solid,$\cos\theta_p = 0.1$; red-dashed, $\cos\theta_p = 1$. Line-sets: lower, $E_p = 500$ MeV; upper, $E_p = 10000$ MeV.}
\caption{Calculated \gray{} albedo spectrum of the Moon. Line-styles: black-solid, total; blue-dotted, limb -- outer $5'$; red-dashed, center -- inner $20'$. Upper solid line: $\Phi = 500$ MV; lower solid line: $\Phi = 1500$ MV. Limb and center components are only shown for $\Phi = 1500$ MV. Data points from the EGRET \citep{Thompson1997} with upper and lower symbols corresponding to periods of lower and higher solar activity, respectively. The differential 1 year sensitivity of the LAT is shown as the shaded region.}
\caption{ The eigenvalue distributions for $M$, \color{red}$U^{(1)}$\color{black}, \color{green}$U^{(2)}$\color{black},..., \color{blue}$U^{(g-1)}$\color{black}, and \color{magenta}$U^{(g)}$\color{black}. The $N$ $m$-eigenvalues split into $g+1$ groups $\{m_i^{(I)}|i=1,...,n_I\}$,~$I=1,...,g+1$. The $k_I$ ($u^{(1)}$-$u^{(2)}$-...-$u^{(I)}$)-bound states are distributed uniformly between $\{m_i^{(I)}\}$ and $\{m_i^{(I+1)}\}$, and are represented by $I$ coincident colored lines in the imaginary direction. }
\caption{ The eigenvalue distributions for $M$, \color{red}$V^{(1)}$\color{black}, \color{green}$V^{(2)}$\color{black},..., \color{blue}$V^{(g-1)}$\color{black}, \color{magenta}$V^{(g)}$\color{black}. The $n_{g+1-I}$ ($m$-$v^{(1)}$-...-$v^{(I)}$)-bound states are distributed according to Wigner's semi-circle law. }
\caption{\label{etamaxfig} (Color) (a) Optimum pulse separation, $\eta_{opt}$ as function of Rabi frequency ratio, $r=\Omega_{A,0}/\Omega_{B,0}$. As indicated, the three curves correspond to different values of $\Lambda$, given by $\Lambda=1$ (\textcolor[rgb]{0,0.6,0}{{\textbf{$\blacklozenge$}}}), $\Lambda=3$ (\textcolor[rgb]{0,0,1}{{\textbf{$\blacktriangle$}}}) and $\Lambda=10$ (\textcolor[rgb]{1,0,0}{{\textbf{$\bullet$}}}). One-photon detunings are $\Delta_{A}/2 \pi=\Delta_{B}/2 \pi= 4$ GHz and all results are obtained with $\tau=2$ $\mu$s. (b) The transfer inefficiency $1-P_{max}$ corresponding to the values of $\eta_{opt}$ calculated in (a). The oscillations of the curves are attributed to Rabi dynamics and it should be noted that the curves are symmetric with respect to letting $r \rightarrow r^{-1}$.}
\caption{(Color) Transfer efficiency as a function of pulse width of the $C$ and $D$ laser fields for $2\pi\times 2$ kHz (\textcolor[rgb]{1,0,0}{\large{\textbf{---}}}) as well as no laser linewidth (\textbf{\large{- - -}}) The delay between the pulses is given by $\eta =0.85$, laser detunings are $\Delta_{C}=\Delta_{D}=2\pi\times 600$ MHz and peak Rabi frequencies $\Omega_{C,0}=\Omega_{D,0}=2\pi\times 100$ MHz ($\Lambda=9.3$).}
\caption{(Color) Optimal pulse width as a function of the Rabi frequencies of the $C$ and $D$ laser fields for $2\pi\times 2$ kHz (\textcolor[rgb]{1,0,0}{{\textbf{$\bullet$}}}) as well as no laser linewidth ({\tiny{\textbf{{$\blacksquare$}}}}). The delay between the pulses is given by $\eta =0.85$, laser detunings are $\Delta_{C}=\Delta_{D}=2\pi\times 600$ MHz.}
\caption{(Color) Transfer efficiency as a function of the micromotion velocity. The different curves show results for different detunings: (\textcolor[rgb]{0,0,1}{\large{\textbf{$\cdot$$\cdot$$\cdot$}}}) $\Delta_A=\Delta_B=$ 0 MHz, (\textbf{\large{---}}) $\Delta_A=\Delta_B= 2\pi\times 300$ MHz $(\Lambda=18.5)$, (\textcolor[rgb]{1,0,0}{\large{\textbf{- - -}}}) $\Delta_A=\Delta_B=2\pi\times 600$ MHz $(\Lambda=9.3)$ and (\textcolor[rgb]{0,1,0}{\large{\textbf{-{$\cdot$}-}}}) $\Delta_A=\Delta_B=2\pi\times 1200$ MHz $(\Lambda=4.6)$. Parameters used for simulations: $\tau_A=\tau_B=2$ $\mu$s, $\Delta t=1.3$ $\mu$s $(\eta=0.92)$, $\Omega_{A}=\Omega_{B}=2\pi\times 100$ MHz and $\Omega_{RF}=2\pi\times 16.8$ MHz. The temperature of the ion is defined as T=$m v^2/k_{B}$, where $m$ is the ion mass and $k_B$ Boltzmann's constant. The Doppler temperature for Ca$^+$, T$_D$, is indicated by the vertical line.}
\caption{(Color) Effect of polarization error and error in magnetic field direction. Populations of the qubit states $\vert\!\!\downarrow \rangle$, $P_-$, and $\vert\!\!\uparrow \rangle$, $P_+$, plotted as a function of the relative $\sigma_-$ polarization component, $\Omega_A ^-/\Omega_A ^+$. The various curves on each plot correspond to different magnetic field direction errors, parameterized by the relative $\pi$ polarization component, $\Omega_A ^0/\Omega_A ^+$. (\textbf{\large{---}}) $\Omega_A ^0/\Omega_A ^+=0$, (\textcolor[rgb]{1,0,0}{\large{\textbf{- - -}}}) $\Omega_A ^0/\Omega_A ^+=0.01$, (\textcolor[rgb]{0,0,1}{\large{\textbf{-{$\cdot$}-{$\cdot$}}}}) $\Omega_A ^0/\Omega_A ^+=0.03$ and (\textcolor[rgb]{0,1,0}{\large{\textbf{$\cdot$$\cdot$$\cdot$}}}) $\Omega_A ^0/\Omega_A ^+=0.05$. The upper two graphs, (a) and (b), correspond to all initial population in the $\vert\!\!\downarrow \rangle$ qubit state, while the lower two graphs, (c) and (d), correspond to all initial population in $\vert\!\!\uparrow \rangle$. The parameters for the simulations are $\Omega_{A,0} ^+ =\Omega_{B,0}=2\pi\times 300$ MHz, $\Delta_A=\Delta_B=2\pi\times 300$ MHz, $\tau=2$ $\mu$s and $\Delta t=1.3$ $\mu$s. ($\Lambda=167$, $\eta=0.92$).}
\caption{(Color) Full simulation of the two stage STIRAP. The plot shows the population transfer as a function of two-photon detuning of the first STIRAP stage. Parameters are: $\Delta_A=\Delta_C=\Delta_D=2\pi\times 600$ MHz, $\Omega_A=\Omega_B=\Omega_C=\Omega_D=2\pi\times 100$ MHz, $\tau_A=\tau_B=\tau_C=\tau_D=2 \mu$s. All laser linewidths are $2\pi\times 2$ kHz and the trap is characterized by $\Omega_{RF}=2\pi\times 16.7$ MHz and $v=0.4$ m/s, corresponding to a temperature of 0.8 mK. The STIRAP's have $\Delta t=1.2 \mu$s corresponding to $\eta=0.85$ and $\Lambda=9.3$ for both stages. The two STIRAP stages are separated by 10 $\mu$s and the overall simulation time is 30 $\mu$s. The Rabi frequencies associated with the residual power are 0.01 relative to the peak Rabi frequencies for all four fields. The curves correspond to: (\textbf{\large{---}}) all effects on, (\textcolor[rgb]{0,1,0}{\large{\textbf{- - -}}}) residual power set to zero but laser linewidth on, and (\textcolor[rgb]{0,0,1}{\large{\textbf{{$\cdot$}{$\cdot$}{$\cdot$}}}}) both laser linewidth and residual power set to zero.}
\caption{(a) The mean number of clicks (\fullsquare) and mean number of photons after deconvolution (\opencircle) are plotted against the transmission. For the deconvoluted data the expected linear fit holds over the full range of transmission, while for the click data non-linear behaviour is visible for higher transmission. The clicks follow \eref{TMD:theorieverlust2}, which can be verified looking at the light grey fitting line. (b) For the deconvoluted data (\opencircle) the Mandel Q-parameter against the transmission shows the expected value of 1 for all transmission levels. For the click data Mandel Q-parameters less than 1 falsely indicate non-Poissonian nature.}
\caption{Evolution of $\langle{\xi^2}\rangle_\pi(\mu^2)$ (left panel) and $\langle{\xi^4}\rangle_\pi(\mu^2)$ (right panel) with $\mu^2\in1-10$~GeV$^2$. The green strip on both panels corresponds to the unified results of the QCD Sum Rules with NLCs (minimal and improved models of the QCD vacuum). On the left panel, we also show the lattice results with their corresponding error-bars: {\small\red\ding{115}} \protect\cite{MS88}, {\small\ding{116}} \protect\cite{DelD05}, {\violet\ding{117}} \protect\cite{Lat06}, {\small\blue\ding{110}} \protect\cite{Lat07}. \label{fig:Mom24.Evol.Lat} }
\caption{Scaling exponent in the subdiffusive region of $J(t)$ for each tracked particle in Fig.~\ref{amoebae}(a), plotted against particle speed (\textcolor{Gray}{$\bullet$}). The data are binned and the mean of each bin plotted as a line graph, with error bars calculated from the standard deviation of each bin. The scaling exponent shows a clear increase with increasing speed.}
\caption{(a) Endoplasmic (\textcolor{Red}{$\bullet$}) and cortical regions (\textcolor{Green}{$\bullet$}) in the lobopod of the large amoeba of Fig.~\ref{amoebae}(b), identified by the direction of particle drift with respect to the direction of the lobopod. (b) Mean particle speed of endoplasmic particles (\Flatsteel), plotted simultaneously with mean subdiffusive scaling exponent ($\dotline$), as a function of time, in steps of 1/3~s. The speed and scaling exponent rise and fall in tandem. (c) Mean particle speed and mean scaling exponent, from (b), are plotted against each other. A clear trend is shown by the linear least-squares fit (\Flatsteel): the data points follow this line much more closely than the error bars which show the standard deviation of the sample of particles at each time point. The trend is significantly shallower here than in the data for the entire small amoeba (\textcolor{Gray}{\Flatsteel}), of Fig.~\ref{smallaprot-shearthinning} above.}
\caption{The mid-section of an \emph{A.~proteus} lobopod is displayed (top), showing the direction of flow. From tracked particles in this image, we calculate a velocity profile (centre) and a scaling exponent profile (bottom). Individual data points (\textcolor{Gray}{$\cdot$}) are plotted on each graph, with the rebinned and averaged data plotted with error bars, showing each standard deviation from the mean (\Flatsteel). The velocity of the endoplasmic layer is fitted well with a parabola ($\dotline$), which corresponds to Poiseuille flow in a Newtonian fluid. The scaling exponent profile shows two regions of scaling exponent $\approx$~0.8 and 1.0, respectively corresponding to the cortex and endoplasm as identified from the velocity profile. The boundary between the layers seems abrupt; there is no gradual change in the mean-scaling-exponent, as far as the local inhomogeneity in the data permits us to discern.}
\caption{Total energies (divided by $4 \pi B$) of the charge--one ({\scriptsize $\fulldiamond$}) charge--two ({\scriptsize \textcolor{Gray}{$\fullsquare$}}) and charge--three ({\scriptsize $\opendiamond$}) Skyrmions as a function of the parameter $s$ for various $\kappa$ values. Each of the energy graphs attains a minimal value at some $s$. At $s \approx 2$ the energy--per--topological--charge of the charge--two and charge--three solutions reaches the charge--one energy (from below), and stable solutions are no longer observed.}
\caption{ Symmetry--breaking measure $\Delta^2$, as a function of $s$ for various $\kappa$ values in the charge--three sector: $\kappa^2=0.01$ ({\scriptsize $\opendiamond$}), $\kappa^2=0.05$ ({\scriptsize \textcolor{Gray}{$\fullsquare$}}), $\kappa^2=0.25$ ({\scriptsize $\fulldiamond$}), and $\kappa^2=1$ ($\times$). Breaking of rotational symmetry becomes more and more apparent as $s$ and $\kappa$ increase. The lines are to guide the eye.}
\caption{\label{replicon} Statistical analysis of the skew profiles of the 287 pairs of \emph{ori} selected as explained in the text. The \emph{ori} spacing $l$ was rescaled to 1 prior to computing the mean $S$ values in windows of width $1/10$, excluding from the analysis the first and last half intervals. (a) Mean $S$ profile ($\bullet$) over windows that are more than 90\% intergenic. (b) Mean $S$ profile ($\bullet$) over windows that are more than 90\% genic; the symbols (\textcolor{gris}{$\blacktriangle$}) (resp. (\textcolor{gris}{$\square$})) correspond to the percentage of sense (antisense) genes located at that position among the 287 putative \emph{ori} pairs. (c) Histogram of the slope $s$ of the skew profiles after rescaling $l$ to 1. (d) Histogram of the mean absolute deviation of the $S$ profiles from a linear profile. \vspace{-0.9cm}\\ }
\caption{Free energy of a single loop of size $l$ in a chain of length $L=10^5$ under the ``winding'' constraint and for $A=200$, $\sigma_o=0.01$ and $\alpha= 2 \pi$. (a) Energy landscape along a chain with uncorrelated ($H=0.5$, black) and LRC ($H=0.8$, grey) disorders. (b) Free energy r.m.s. fluctuations $\Lambda(l)$ {\em vs.} $l$; the symbols correspond to numerical estimate for different disorders: $H=0.5$ ($\circ$), $0.6$ ($\square$), $0.7$ (\Pisymbol{pzd}{73}), $0.8$ ($\triangle$) and $0.9$ (\small$\Diamond$\normalsize); the solid lines correspond to Eq.~(\ref{fm}). (c) Reduced correlation function $C(y,200)/\Lambda(200)$ {\em vs.} $y$; the symbols have the same meaning as in (b); the solid curves correspond to Eq. (\ref{eq_corr}).}
\caption{Free energy of the single loop system $\mathcal{F}_H(l,L)$ {\em vs} $l$ under the ``cyclization'' constraint and for $L=15000,~A=200,~\alpha=2 \pi$ and $\sigma_o=0.01$ (a) and $0.05$ (b). The symbols correspond to a single chain's free energy $\mathcal{F}_H(l,L)$ obtained from the (exact) numerical computation of $f(s,l)$ for $H=0.5$ ($\circ,~\bullet$), $0.6$ (\small $\square,~\blacksquare$ \normalsize), $0.7$ (\small \Pisymbol{pzd}{73},\small \Pisymbol{pzd}{72}), $0.8$ ($\vartriangle,~\blacktriangle$) and $0.9$ (\small $\Diamond$, \ding{117}); the ($\times$) correspond to the ``pure'' case without disorder. The continuous curves stand for the corresponding quenched free energies $\overline{\mathcal{F}}_H(l,L)$ averaged over $100$ ``typical'' single chain free energies computed using Eq.~(\ref{eq_enth}) for the enthalpic part ({\em i.e.} the ``winding'' energy) and $c_c= 7/2$ for the entropy cost; the dotted curve correspond to the exact analytical expression for the ``pure'' case. (c) $\mathcal{F}_{H}(l)-\mathcal{F}_{1/2}(l)$ {\em vs} $H$ between LRC and uncorrelated chains, for loop size $l=200$; the dashed curves correspond to perturbative approximation. (d) Optimal loop length $l^{\ast}(H)$ {\em vs} $H$; the dashed curves correspond to the perturbative expression (\ref{eq_lopt}); the horizontal line indicates the optimal loop length $l^*=1128$ for the ``pure'' system. In (c) and (d) the symbols and the continuous curves have the same meaning as in (a) and (b)}
\caption{\label{fig1}Left: the mid-rapidity m$_t$-spectra for $^3$He (upper panel) and $t$ (lower panel) for the 7\% most central Pb+Pb collisions (dashed lines show the double-exponential fits used for extrapolation to the unmeasured range). Right: (upper panel) $<$$m_t$$>$-m versus particle mass as obtained from the fits to the spectra at 20A(\fullsquare) and 80A GeV(\fullcircle); (lower panel) mid-rapidity $t$ to $^3$He ratio as measured (points) and predicted by the SHM model \cite{becat} (band).}
\caption{\label{fig3}Left: R$_{coal}$ for $d$(\fullsquare) and $^3$He(\fullcircle) in central A-A collisions at AGS (blue), SPS (red) and RHIC (green). Center: invariant cross section per wounded nucleon at $p_t$=0 as a function of $<$N$_w$$>$ for anti-protons (\fullcircle) and anti-deuterons (\fullsquare). Right: B$_2$ for deuterons (\opensquare) and anti-deuterons (\fullsquare) as a function of $<$N$_w$$>$.}
\caption{\label{fig:gamma} (Color) Susceptibility and Specific heat. (a) Temperature $T$ dependence of real part of AC susceptibility, $\chi'$. Inset: $\chi'(H)$ at 4 K with power law fit (red line). (b) $T$ dependence of $1/\chi'$ at 0 and 1 mT. Red line is fit of Curie-Weiss form at high-$T$. Inset: $1/\chi'(0)$ vs.\reduced$T$, ($T/T_C^{r}-1$). Red line is $1/\chi'=a(T-T_C^{r})^{\delta}$ with $\delta = 0.45\pm 0.04$ and $T_C^{r}=24 \pm 0.5$ K. (c) Specific heat, $C$ divided by $T$, vs.\$T$ for $x=0.002$ (circles), $0.005$ (bullets), $0.007$ at $H=0$ (blue squares), $H=1$ T (solid line), and $H=3$ T (dashed line), $0.03$ (FM) (triangles), and $0.045$ (FM) (diamonds). Red line is fit of the form $aT^{-\alpha}$ with $\alpha = 0.69\pm 0.05$, for $x=0.005$. Dotted lines are fits of a Sommerfeld plus Schottky model to the data at $T > 2$ K.}
\caption{Maximal Sweeps on \triangleboard$(n)$, where $n=3$, $5$, $7$ and $9$. These sweeps are shown starting and ending at a1, but can begin and end at other board locations.}
\caption{A 15-move solution to problem \#1 on\triangleboard$(8)$ (finishing with an 18-sweep). Note: more than one move is sometimes shown between board snapshots.}
\caption{Fitting the inductive components together to make a long finishing sweep on \triangleboard$(24)$.}
\caption{Statistics on long sweep solutions on \triangleboard$(12i)$.}
\caption{The minimal 13-move solution to the a7-complement on \triangleboard$(8)$. Note: more than one move is sometimes shown between board snapshots.}
\caption{Dividing \triangleboard$(6)$, \triangleboard$(8)$ and \triangleboard$(10)$ into ``Merson Regions" ($R$ is the number of regions).}
\caption{Managing a session in a thread. The \index{reply/1}\predref{reply}{1} predicate is part of the HTTP worker pool, while \index{session_loop/1}\predref{session_loop}{1} is executed in the thread handling the session. We omitted error handling for readability of the example.}
\caption{Loading triples using \index{process_rdf/3}\predref{process_rdf}{3}}
\caption{Split rdf conjunctions. After executing the first \index{rdf/3}\predref{rdf}{3} query \arg{Author} is bound and the two subsequent queries become independent. This is also true for other orderings, so we only need to evaluate 3 alternatives instead of 3! (6).}
\caption{Experimental results for the cat state. {\color{Red}{$% \blacktriangledown$}} NMR experiment, {\color{Blue}{$\blacksquare$}} Photon experiment taken from \protect\cite{aspect}. The solid line is the Quantum Mechanical predictions.}
\caption{Experimental results of the $CHSH$ quantity as function of the angle $\protect\theta $. (a) {\color{Red}{$\blacktriangledown$}} $\left\vert 00\right\rangle $, {\color{Blue}{$\blacksquare$}} $(\left\vert 00\right\rangle +\left\vert 01\right\rangle +\left\vert 10\right\rangle +\left\vert 11\right\rangle)/2$. (b) {\color{Red}{$\blacktriangledown$}} $(\left\vert (00\right\rangle + \left\vert 11\right\rangle)/\protect\sqrt{2} $, {\\color{Blue}{$\blacksquare$}} $(\left\vert 01\right\rangle - \left\vert 10\right\rangle)/\protect\sqrt{2} $. The continuous lines are the predictions of the LRHVM described in \protect\cite{menicucci}. The NMR data showed here are the same as in figure (\ref{nmrphoton})}
\caption{Maximal sweeps on the standard boards, only the second is a super-sweep. The special green (or shaded) peg performs the sweep. (a) A 16-sweep on the 33-hole cross-shaped board. (b) A 9-sweep on the 15-hole triangular board, \triangleboard$(5)$.}
\caption{Colors indicate the length of the shortest solution to a complement problem on \triangleboard$(8)$ and \rhombusboard$(6)$ (magenta = 13 moves, cyan = 14 moves, yellow = 15 moves). The \rhombusboard$(6)$ board has been rotated to its ``diamond" configuration to show the symmetry.}
\caption{For comparison with the simulation results of Fig.~\ref{fig:Birdfoot_simu1} the figure shows part of a map of the mouth of the Mississippi river, where the birdfoot shaped delta can be seen clearly. The colors indicate channel deposits \colorbox{ChannelDeposit}{\textcolor{ChannelDeposit}{x}}, sand ridges \colorbox{SandRidges}{\textcolor{SandRidges}{x}}, swamps \colorbox{Swamp}{\textcolor{Swamp}{x}} and marshes \colorbox{Marsh}{\textcolor{Marsh}{x}}. The figure was generated after \cite{Coleman76}.}
\caption{In (a) the simulation of a wave-dominated delta is shown. While the waves are reworking the coast at the mouth of the river to form an estuary, the river deposits sediment and forms large beaches. As the simulation does not include evaporation, the ponds and inactive channels in the deposition zone do not disappear as in the map of the real river shown in (b). The parameters in the simulation were $N=179$, $I_0=1.7\times 10^{-4}$, $s_0=0.0015$, $c_\sigma=8.5$, $c_1=0$, $c_2=0.1$ and $I^\star=1.3\times 10^{-4}$. Smoothening was applied every 200 time steps with a smoothening constant $\epsilon=0.01$. In (b) we show for comparison a map of the S\~{a}o Francisco river delta in southern Brazil which is the most wave-dominated delta according to the classification of \cite{Galloway75}. The colors in the map (Fig.~(b)) indicate channel deposits \colorbox{ChannelDeposit}{\textcolor{ChannelDeposit}{x}}, beach ridges \colorbox{BeachRidges}{\textcolor{BeachRidges}{x}}, eolian dunes \colorbox{EolianDunes}{\textcolor{EolianDunes}{x}}, marsh-mangroves \colorbox{Marsh-mangroves}{\textcolor{Marsh-mangroves}{x}}, the floodplain \colorbox{Floodplain}{\textcolor{Floodplain}{x}} and the uplands \colorbox{Uplands}{\textcolor{Uplands}{x}}. The figure was generated after \cite{Coleman76}. }
\caption[ Temperature dependence of the unit volume free energy of cavitation for small repulsive spheres of growing size. Small-sized cavities of radii B = 2 \AA \, (squares) and B = 3\AA \, (discs) exhibit room temperature tangents with positive slope that correspond to a unit volume$\Delta$S of negative sign. Similar analysis of tangential slopes for larger volumes, i.e. B = 4 \AA \, (triangles) and B = 5\AA \, (diamonds) demonstrates that the magnitude of the slopes decreases at 300 K when the size of the solute becomes bigger.]{ \label{fig1} Temperature dependence of the unit volume free energy of cavitation for small repulsive spheres of growing size. Small-sized cavities of radii B = 2 \AA \, (squares) and B = 3\AA \, (discs) exhibit room temperature tangents with positive slope that correspond to a unit volume$\Delta$S of negative sign. Similar analysis of tangential slopes for larger volumes, i.e. B = 4 \AA \, (triangles) and B = 5\AA \, (diamonds) demonstrates that the magnitude of the slopes decreases at 300 K when the size of the solute becomes bigger.}
\caption[ Temperature dependence of the full size free energy of cavitation for small repulsive spheres of growing size. Room temperature tangents exhibit positive slopes that correspond to a $\Delta$S of cavitation of the entire cavity volume with negative sign. The magnitude of the slope of tangential lines increases with growing cavity size, i.e. when following cavities of radii B = 2 \AA \, (squares), B = 3\AA \, (discs), B = 4\AA \, (triangles) and B = 5\AA \, (diamonds). Inversion in the sign of the slopes of tangential lines is seen close to the boiling point.]{ \label{fig2} Temperature dependence of the full size free energy of cavitation for small repulsive spheres of growing size. Room temperature tangents exhibit positive slopes that correspond to a $\Delta$S of cavitation of the entire cavity volume with negative sign. The magnitude of the slope of tangential lines increases with growing cavity size, i.e. when following cavities of radii B = 2 \AA \, (squares), B = 3\AA \, (discs), B = 4\AA \, (triangles) and B = 5\AA \, (diamonds). Inversion in the sign of the slopes of tangential lines is seen close to the boiling point.}
\caption[ Comparison of simulated and measured surface tension, $\sigma$, of liquid water at various temperatures. The calculated values of $\sigma$ (squares) are due to $\lim\limits_{{\rm B} \to \infty}\Delta$ G$^{cav}/\textcolor{red}{(}4\pi$ B$^2$\textcolor{red}{)} with B=100 \AA. Experimental reference data (discs) have been obtained from standard tabulations \cite{iapws}. Although qualitatively comparable, the calculated values of $\sigma$ exhibit inverse temperature profiles, which might be an artefact stemming from large-scale extrapolation. ]{ \label{fig4} Comparison of simulated and measured surface tension, $\sigma$, of liquid water at various temperatures. The calculated values of $\sigma$ (squares) are due to $\lim\limits_{{\rm B} \to \infty}\Delta$ G$^{cav}/\textcolor{red}{(}4\pi$ B$^2$\textcolor{red}{)} with B=100 \AA. Experimental reference data (discs) have been obtained from standard tabulations \cite{iapws}. Although qualitatively comparable, the calculated values of $\sigma$ exhibit inverse temperature profiles, which might be an artefact stemming from large-scale extrapolation. }
\caption{ \coloronline In \subref{fig:Honeycomb} $\vec{a}_1$ and $\vec{a}_2$ are the primitive vectors that define the \ac{WS} unit cell highlighted as the dashed hexagon. The lattice parameter, $a$, is $\simeq 1.4$ \AA. The first \ac{BZ} of the associated reciprocal lattice is shown in (b), together with the points of high symmetry $\Gamma$, $M$ and the two nonequivalent $K$ and $K'$. }
\caption{ \coloronline \subref{fig:BandSelected} Band structure along the symmetry directions of the reciprocal \ac{BZ} of the honeycomb lattice. \subref{fig:HoneycombSpectrum} Band structure of graphene ($t'=0$) with the two bands touching at the $K$ and $K'$ points of the \ac{BZ}. }
\caption{ \coloronline \ac{DOS} associated with eq.~\eqref{eq:Dispersion} for different values of the next nearest neighbor hopping $t'$. }
\caption{ \coloronline Selected eigenstates in a graphene sheet with $80^2$ atoms containing a single impurity at the center (black dot). Only the region near the vacancy is shown. \subref{fig:Eigenstate-NearZero} The eigenstate with energy closest, but different, to zero.. \subref{fig:Eigenstate-AtZero} The eigenstate with $E=0$. \subref{fig:Eigenstate-Tlinha-1} and \subref{fig:Eigenstate-Tlinha-2} show the presence of two quasi-localized eigenstates even with $t=0.2t$. }
\caption{ \coloronline \ac{IPR} and \ac{LDOS} calculated at one site closest to the vacancy. In panel \subref{fig:PR-1}, we have results for the \ac{IPR} with $t'=0$ and $t'\ne 0$ without any vacancy (top row), and with a single vacancy (bottom) for comparison. In panel \subref{fig:PR-2} we show the dependence of the \ac{IPR} of the zero mode, $\mathcal P(E=0)$, with the system size $N$ (left), and also $\average{\mathcal P(E)}$ versus $N$ for the remainder (extended) states (right). Dashed lines are guides for the eye. }
\caption{ \coloronline \ac{IPR} and \ac{DOS} for the diluted honeycomb lattice. \subref{fig:Dilution-LDOS} \ac{DOS} for selected concentrations, $x$, and different values of $t'$. \subref{fig:Dilution-PR} \ac{IPR} for selected values of $t'$ using a concentration $x=0.5\,\%$. For comparison, the corresponding \ac{DOS} is also plotted in each case. The concentration of vacancies is $x$, and only the vicinity of the Fermi level is shown. }
\caption{ \coloronline Dilution of just one sublattice of the honeycomb. \subref{fig:SameSublattDilution-1} \ac{DOS} for different dilution strengths, diluting only sublattice $A$. \subref{fig:SameSublattDilution-2} On the left we show a detail of the \ac{DOS} and the evolution of the gap with vacancy concentrations. On the right we plot the dependence of the missing spectral weight on the band ($=1-w_\delta$) with $x$ (circles). The continuous line is the best fit using $f(x) = a x /(b - x)$ to the data represented by the circles. }
\caption{ \coloronline The gap estimated from the numerical curves in Fig.~\ref{fig:SameSublattDilution} is plotted against the vacancy concentration, $x$. The continuous line is a least squares fit to $f(x) = ax^b$. }
\caption{ \coloronline \ac{DOS} for the honeycomb lattice using the controlled selective dilution discussed in the text, calculated for different concentrations of vacancies, $x$, and several degrees of uncompensation, $\eta$. Only the low energy region close to the Dirac point is shown. }
\caption{ \coloronline Real part of \eqref{eq:KramersKronig} (top) obtained from the homogeneous \ac{DOS} (bottom) through the Kramers-Kronig relations. }
\caption{ \coloronline (Top) \ac{LDOS} at the impurity site ($\rho_0(E)$) for different strengths of the scattering potential, indicated in the legend. (Bottom) The same data divided by the free \ac{DOS} ($\rho^0(E)$). }
\caption{ \coloronline \subref{fig:Binnary-Eres} Position of the maximum in the \ac{LDOS} compared with the roots of \eqref{eq:ResonanceCondition}. \subref{fig:Binnary-Ebound} Energy of the bound state. }
\caption{ \coloronline DOS of the honeycomb lattice with a finite density of local impurities. \subref{fig:Binnary-Ea-12} shows the \ac{DOS} for $U=12t$ and different concentrations of impurities (notice the truncation in the horizontal axis). \subref{fig:Binnary-DOS-vs-Ea-and-x} shows a detail of the low energy region for different $U$ and $x$ as noted in the different graphs. }
\caption{ \coloronline \subref{fig:Binnary-DiracPt-vs-x-and-U} Variation of the \emph{``Dirac point''} energy $E_D$ with impurity concentration and strength. The inset shows $E_D/EU$ as a function of $U$, in which the curves with $U\le 3$ roughly collapse onto each other. \subref{fig:Binnary-SW-ImpBand} Spectral weight transfer to the impurity band in the presence of local impurities. The values shown correspond to integration of the global \ac{DOS} beyond $E=4 t$ (above the main band edge, cfr. Fig.~\ref{fig:Binnary}). }
\caption{ \coloronline Effect of a single substitutional impurity in the \ac{LDOS}. In panel \subref{fig:Hopping-Pert-LDOS1} we plot the \ac{LDOS} calculated at the site of the impurity for the four different values of $t_0$ indicated in each frame. In \subref{fig:Hopping-Pert-LDOS2} the situation is identical but the \ac{LDOS} is calculated at the nearest neighboring site of the impurity. }
\caption{ \coloronline Variation of the energy corresponding to the peak in the \ac{LDOS} with the magnitude of $t_0$. The \ac{LDOS} in question is the \ac{LDOS} calculated at the impurity site. }
\caption{ \coloronline The \ac{DOS} corresponding to the model Hamiltonian \eqref{eq:Hamiltonian-HoppingPert}, with a finite density of impurities. The three panels correspond to different values of the perturbing hopping ($t_0=0.5$, $0.8$, $0.9$ and $0.95$), and within each panel the three curves were obtained at different concentrations ($x=0.01$, $0.05$ and $0.1$). The inset of the bottom panels is a magnification of the region near $E=0$. }
\caption{Calculated \gray{} albedo spectrum for CR nuclei interactions in the Moon rock \cite{MP2007b} for selected modulation potentials. Line colouring: black, no modulation; red, $\Phi = 500$ MV; blue, $\Phi = 1500$ MV. Dashed and dotted lines show the albedo of the disk and the rim correspondingly for the case of $\Phi=1500$ MV. }
\caption{Calculated \gray{} albedo spectrum of a Moon-sized body at the Lunar distance composed of moon rock (black), iron ($\times10$, blue), or water ice ($\times0.1$, red). Line-styles: solid, no modulation; long-dashed, $\Phi = 1500$ MV.}
\caption{Calculated \gray{} albedo spectrum showing components below 20 MeV for no modulation (red) and modulation level 1500 MV (blue). Line-styles: long dash: positron induced \gray{s} from center; short dash: CR positron induced \gray{s} from rim; thin solid: total CR positron induced \gray{s}; thick solid: total \gray{} emission from CR positrons and nucleons. }
\caption{ Profiles of \gray{} intensity with $\beta$ derived from EGRET data as described in the text. The energy range is 100--500 MeV and the profiles have been averaged over all ecliptic longitudes. $(a)$ Profile derived with no masking of Galactic diffuse emission or \gray{} point sources. $(b)$ Profile with the Galactic plane ($|b| < 10^\circ$ for $|l| > 90^\circ$ and $|b| < 20^\circ$ for $|l| < 90^\circ$) excluded. $(c)$ Profile with the identified 3EG sources \citep{Hartman1999} and the Galactic plane excluded. $(d)$ Profile with the identified 3EG sources plus the further blazar identifications proposed by \citet{Sowards2003,Sowards2004} excluded. Overlaid on each profile is the best-fitting gaussian (12.5$^\circ$ FWHM, centered on $\beta$ = 0) plus a constant, fit for the region $|\beta| < 50^\circ$. This approximates the distribution of albedo \gray{} emission expected for the the KBO. }
\caption{Calculated \gray{} albedo spectrum for CR nuclei interactions in the Moon rock \cite{MP2007b} for selected modulation potentials. Line colouring: black, no modulation; red, $\Phi = 500$ MV; blue, $\Phi = 1500$ MV. Dashed and dotted lines show the albedo of the disk and the rim correspondingly for the case of $\Phi=1500$ MV. }
\caption{Calculated \gray{} albedo spectrum of a Moon-sized body at the Lunar distance composed of moon rock (black), iron ($\times10$, blue), or water ice ($\times0.1$, red). Line-styles: solid, no modulation; long-dashed, $\Phi = 1500$ MV.}
\caption{Calculated \gray{} albedo spectrum showing components below 20 MeV for no modulation (red) and modulation level 1500 MV (blue). Line-styles: long dash: positron induced \gray{s} from center; short dash: CR positron induced \gray{s} from rim; thin solid: total CR positron induced \gray{s}; thick solid: total \gray{} emission from CR positrons and nucleons. }
\caption{ Profiles of \gray{} intensity with $\beta$ derived from EGRET data as described in the text. The energy range is 100--500 MeV and the profiles have been averaged over all ecliptic longitudes. $(a)$ Profile derived with no masking of Galactic diffuse emission or \gray{} point sources. $(b)$ Profile with the Galactic plane ($|b| < 10^\circ$ for $|l| > 90^\circ$ and $|b| < 20^\circ$ for $|l| < 90^\circ$) excluded. $(c)$ Profile with the identified 3EG sources \citep{Hartman1999} and the Galactic plane excluded. $(d)$ Profile with the identified 3EG sources plus the further blazar identifications proposed by \citet{Sowards2003,Sowards2004} excluded. Overlaid on each profile is the best-fitting gaussian (12.5$^\circ$ FWHM, centered on $\beta$ = 0) plus a constant, fit for the region $|\beta| < 50^\circ$. This approximates the distribution of albedo \gray{} emission expected for the the KBO. }
\caption{Additional \aastex\symbols}