dedup-isc-ft-v107-score
float64 0.32
1
| uid
stringlengths 32
32
| text
stringlengths 0
32.5k
| paper_id
stringlengths 1
14
| original_image_filename
stringlengths 5
222
|
---|---|---|---|---|
0.522296 | 280a16cac8d447feb8154e44bc3e19a2 | Lensing probabilities and image multiplicities with triaxial dark halos as a function of image separation θ. The source is placed at z_ S=2.0, and slope of the source luminosity function is fixed to β=2.5. | astro | ph0408573-lat_sep.jpg |
0.523122 | 0c73f3b63580409391b101c7d141a528 | Differential distributions of image separation distributions (eq. [<ref>]). We fix the source at z=2 and adopt a power-law luminosity function ϕ_L∝ L^-2.5. We use α=1 for the fiducial value. Dependences of several parameters are shown; Ω_M (upper left), σ_8 (upper right), and α (lower left). It is also demonstrated how distributions change by adopting different models of concentration parameters (lower right). | astro | ph0408573-lat_tdis.jpg |
0.478042 | 800813cf8f594c339b9a6dc43c1ecb9e | Comparison of two timescales, t_ cool (eq. [<ref>]) and t_ age (eq. [<ref>]), as a function of the virial mass of dark halos M_ vir. Two timescales become equal at M_ vir=M_ cool∼ 10^13h^-1M_⊙. | astro | ph0408573-lat_time.jpg |
0.532759 | 182bab93a9374d13a593dec6537d1d8b | Lensing probabilities and image multiplicities as a function of source redshift z_ S. We fix the image separation to θ=15” and the of the source luminosity function to β=2.5. | astro | ph0408573-lat_zs.jpg |
0.529733 | 5d6d9853fa264296a0991b652ef07b90 | The schematic meaning of κ, γ_1, and γ_2 in the Jacobi matrix (<ref>). The convergence κ expands a light bundle, while the shear γ changes the shape. | astro | ph0408573-lens_convshear.jpg |
0.529897 | 43771b6867364f7ba54bddd7d3c84f71 | Vector fields around three stationary point of ϕ. | astro | ph0408573-lens_index.jpg |
0.448959 | d98130e8bd1f43229ae7820f2f9f0c50 | A schematic diagram of the lensing system. The image and source positions in each plane are denoted by ξ⃗ and η⃗. Angular diameter distances should be used for the distances shown in this Figure. | astro | ph0408573-lens_lenseq.jpg |
0.435108 | 152e679846184f22aefa635db8f169e1 | Calculation of the geometrical time delay. We consider the plane on which the source, observer, and deflection point (image) exist. From the purely geometrical consideration, one finds Δℓ≃βζ and β=D_ OSα̂/2D_ LS. | astro | ph0408573-lens_lenseq_td.jpg |
0.52069 | 1b06a5b599fd4d75af59b13d5e1d4461 | Power spectra for three dark matter models; CDM, WDM, and HDM. | astro | ph0408573-pk_dm.jpg |
0.40118 | ace2cf7a82244008b7966cdc688dccf9 | Comparison of transfer functions (eq. [<ref>]) for different values of Ω_M and h (we fix Ω_bh^2=0.022). We compare transfer functions obtained from the Boltzmann equation (Full), fitting formula of <cit.> including baryon wiggle (EH), fitting formula of <cit.> without baryon wiggle (EH(nowig)), and fitting formula of <cit.> with a shape parameter of equation (<ref>) (BBKS). For each set of cosmological parameters, we show transfer functions (upper) and the fractional deviations from the “Full” transfer function (lower). | astro | ph0408573-pk_trans.jpg |
0.412216 | 518dd9dacf7e495294560cbd9c3a972b | Distributions of galaxies brighter than i=24 with (right) and without (left) the color cut. The origin (0,0) is set to the position of the central galaxy G1. Filled squares denote the four lensed images. | astro | ph0408573-sdss_colorcut.jpg |
0.454726 | a1b9950f7a2c4010bc48ca139a0ffeaa | Color-magnitude diagrams for the SDSS J1004+4112 field taken with Suprime-Cam. We divide the galaxies into three categories according to their positions: filled circles denote galaxies inside a 40”×40” box centered on G1, open triangles denote galaxies inside a 100”×100” box, and crosses denote galaxies inside a 200”×200” box. These box sizes correspond to 0.2h^-1 Mpc×0.2h^-1 Mpc, 0.5h^-1 Mpc×0.5h^-1 Mpc, and 1.0h^-1 Mpc×1.0h^-1 Mpc at z=0.68, respectively. Three spectroscopically confirmed member galaxies are marked with open circles. The corresponding r-band absolute magnitudes at z=0.68 (without K-correction) are given at the top of the frame. | astro | ph0408573-sdss_colormag.jpg |
0.471396 | c8ea2d0fdcb74727ba3e1541edad8920 | The cross-filter K-correction, computed from the SDSS composite quasar spectrum created by <cit.>. Dotted line indicate the approximation (eq. [<ref>]) with α_ s=0.5. | astro | ph0408573-sdss_kcor.jpg |
0.450659 | 30195edbc7d1418a857b7b15e1badc95 | Seven epoch data of C4 emission lines of components A and B. A power-law continuum is subtracted from each spectrum. The spectra are normalized to the peak of C4 emission lines. See <cit.> for more details. | astro | ph0408573-sdss_micro.jpg |
0.513655 | 500a3a3b5d4041029974d1438095efc6 | Maximum likelihood estimates for σ_8 with the triaxial model, obtained by combining the discovery of SDSS J1004+4112 in SDSS with the lack of large-separation lenses in CLASS. In making predictions for SDSS, we consider two cases: the appropriate prediction could be the total number of lenses (solid); or since SDSS J1004+4112 is a quad the appropriate quantity could be the number of quadruples (dashed). The likelihoods for α=1 and 1.5 are shown by thick and thin lines, respectively. Results for spherical halos are also shown by dash-dotted lines for reference. We note that this time we used c_ JS as a model for the concentration parameter, and we adopt several approximations, thus the result with the spherical model is slightly different from the result in <ref>. | astro | ph0408573-sdss_s8.jpg |
0.471602 | 56f2cd4dc8cd4b38a812fbad4ca2d5ff | SDSS i^*-band image of SDSS J1004+4112. Components A, B, C, and D are lensed images while component G1 is the brightest galaxy in the lensing cluster. | astro | ph0408573-sdss_sdss1004_i.jpg |
0.533741 | 0f815d50f3d142e7b2c95ee6431da4e1 | Lensing probabilities and image multiplicities for SDSS quasars at redshifts 0.6<z_ S<2.3. We adopted the triaxial lens model to compute the probabilities, which was described in Chapter <ref>. | astro | ph0408573-sdss_sdss_sep.jpg |
0.487134 | ebb2a6fc8c7e410a8d8e6e74950e3034 | Number distribution of strongly lensed quasars known so far. No confirmed lens system has image separation θ>7”. The data are taken from <cit.>. | astro | ph0408573-sdss_sep.jpg |
0.480882 | 59c9521e0f3f4e279a55b19537edb7c6 | Spectra (top) and flux ratios (bottom) of SDSS J1004+4112 components A, B, C, and D taken with LRIS on Keck I. In the upper panel, we can confirm that all components have Lyα, Si4, C4, C3], and Mg2 emission lines at z=1.734. The flux ratios shown in the lower panel are almost constant for a wide range of wavelength. Several absorption lines are also seen in the spectra (see text for details). | astro | ph0408573-sdss_spec.jpg |
0.471418 | d3b420da026548e4ab0fa56eabe03e36 | Spectra of galaxies G2 and G3 taken with FOCAS on the Subaru 8.2-m telescope. From the absorption lines Ca2 H&K, Hδ, and G-band, we find that the redshifts of both galaxies are z=0.675 (z=0.6751±0.0001 from the Hδ lines). | astro | ph0408573-sdss_spec_focas.jpg |
0.439704 | 1c5581f115a647dd93585b746de86306 | Spectrum of the galaxy G1 taken with LRIS on Keck I. The break, Ca2 H&K absorption lines, and Mg absorption line are consistent with redshift z=0.680 (z=0.6799±0.0001 from the Ca2 H line). The G-band also appears in the spectrum. | astro | ph0408573-sdss_spec_g.jpg |
0.464289 | 60b135b7123748c98c1defeefab89bc3 | The gri composite Subaru image of the field around SDSS J1004+4112. Many faint galaxies can be seen — their positions and colors are consistent with being members of a cluster (z=0.68) centered on component G1. | astro | ph0408573-sdss_subaru1004color.jpg |
0.424199 | 8a15dbcc059d4ed7a27901bece445efe | The central region of the Suprime-Cam i-band image. The galaxies with measured redshifts (G1 from LRIS and G2 and G3 from FOCAS) as well as the four lensed images are labeled. The possible lensed arclets are marked with rectangles. | astro | ph0408573-sdss_subaru_center.jpg |
0.445285 | beda40b8832545749dee6e71a11f2f66 | Redshift distribution of quasars identified by the spectroscopic pipeline in the SDSS. Dashed vertical lines show the redshift cut 0.6<z<2.3 used for the statistical analysis. | astro | ph0408573-sdss_z_qso.jpg |
0.4571 | fafeac3c07bb4d8ca9b0936cb78b69da | Dependence of the concentration parameter c_e on the axis ratio. The fitting formula (prefactor in equation [<ref>]) is also shown by the solid line. This Figure is taken from JS02. | astro | ph0408573-tri_c_ac.jpg |
0.458935 | b1eb7c44aec94e05be7ea4319048db54 | Illustration on how critical curves (upper) and caustics (lower) changes as the ellipticity of the projected mass density distribution is increased (from e=0 to e=0.8). In this plot, we adopt the elliptical NFW density profile. The first column shows those of the spherical mass distribution. | astro | ph0408573-tri_crit_e.jpg |
0.416954 | ea51a9bbf3c54f74b39b18701f203026 | Distributions of cross sections for multiple images (defined in the lens plane), both for α=1 and 1.5. We consider a halo with mass M_ vir=10^15h^-1M_⊙ and redshift z=0.3, and the source plane is placed at z=2.0. We took account of PDFs of a/c (eq. [<ref>]), a/b (eq. [<ref>]), and c_e (eq. [<ref>]). We assume that the orientations of dark halos are random. Distributions of cross sections for the corresponding spherical dark halo is also shown for reference. | astro | ph0408573-tri_csdis.jpg |
0.527327 | 5fe273721e3f48f0a4b1da1ff4f5d7c0 | Comparison of f_ GNFW(r) for α=1.5 (denoted by “Integral”; see eq. [<ref>]) and its fitting formula (denoted by “Fit”; see eq. [<ref>]). “Error” plotted in the bottom panel is defined by [f_ GNFW( Integral)-f_ GNFW( Fit)]/f_ GNFW( Integral). | astro | ph0408573-tri_fit.jpg |
0.430988 | 987b3b99f8c2469f96023ca8b10574af | Examples of projected particles distributions in a cluster-size halo. From top to bottom, particles in the isodensity shells with A=100, 2500, 6.25× 10^4 are plotted. The bottom panels show triaxial fits to five isodensity surfaces defined in equation (<ref>). This Figure is taken from JS02. | astro | ph0408573-tri_iso.jpg |
0.444877 | e6176d1efafd4bcc985b1b48c19b5015 | PDFs of q (eq. [<ref>]), q_x (eq. [<ref>]), and q_y (eq. [<ref>]). Here we consider a halo with mass M_ vir=10^15h^-1M_⊙ and redshift z=0.3, but the result only weakly depends on the halo mass and redshift. These PDFs are calculated from PDFs of axis ratios p(a/c) and p(a/b) for which we use equations (<ref>) and (<ref>). We assume that the orientations of dark halos are random. | astro | ph0408573-tri_probq.jpg |
0.510327 | e28c9ba98f0d4b01bd229b7d55ae89b8 | Critical curves (upper) and caustics (lower) of triaxial dark halos projected along three principal vectors (see Figure <ref>). We consider a halo with mass M_ vir=10^15h^-1M_⊙ and redshift z=0.3, and the source plane is placed at z=2.0. The axis ratios are fixed to a/c=0.5 and b/c=0.7. For the concentration parameters, we use the median value given by equation (<ref>). The size of all boxes is the same. The inner slope is α=1. The case for the corresponding spherical dark halo is also shown for reference. | astro | ph0408573-tri_proj.jpg |
0.464005 | d97aeeb6d27d46529a4668cbfd4dbdc4 | The orientations of the coordinate systems. The Cartesian axes (x,y,z) represent the halo principal coordinate system while the axes (x',y',z') stand for the observers coordinate system with z'-axis aligned with the line-of-sight direction. The x'-axis lies in the x-y plane. The angle (θ,ϕ) represent the polar angle of the line-of-sight direction in the (x,y,z)-coordinate system. | astro | ph0408573-tri_shape.jpg |
0.419558 | 9549df3066634e04aeb1cde1d403278d | An histogram of the Hipparcos parallaxes for the Pleiades done with VOPlot, the graphical plug-in of the AVO prototype. Note the peak at around 8 - 9 mas, consistent with the cluster value of 8.46±0.22 mas, and the presence of many foreground and some background stars. | astro | ph0408575-JIL2_Fig1.jpg |
0.443812 | 3bb7d9c63ff34eca9020954593ce7dd9 | The cross-match plug-in of the AVO prototype, used here to cross-correlate the <cit.> catalogue of chemically peculiar stars with WGACAT, a catalogue of all ROSAT observations. | astro | ph0408575-JIL2_Fig4.jpg |
0.412215 | 0ce3ded892cb49de84d94525ff347f8c | The energy of a phase slip center, numercailly interated from Eq.(<ref>) for a third order phase transition free energy. | cond | mat0408002-e3_g.jpg |
0.445989 | a57e55a20b394b649308711fa7a9a520 | The function f(y) as a function of y for a third order free energy. | cond | mat0408002-f3.jpg |
0.453859 | 161c84e0b18e4401be6656ae87e81261 | Order parameter profile f_3 (y) for a third order phase transition in the presence of a current with g = 0.035, 0.134 and 0.250. | cond | mat0408002-f3_g.jpg |
0.490554 | 076c332bfdf34ec79a72a093738ce7b9 | The approximate solution, Eq. (<ref>) (solid line), and numerically exact solution (crosses) of the function f(y)=n^3 for a fourth order phase transition, obtained from Eq.(<ref> | cond | mat0408002-f41.jpg |
0.451238 | c9b3cec5cd1644e5831566706af31703 | Domain wall profile n(y) for a third order phase transition. | cond | mat0408002-n3.jpg |
0.461991 | b16832803a6b4c6fac98c5d8930f2d4b | The magnetization, M, as a function of Zeeman magnetic field for F_2 N_0=0.1 and μ = - 0.16. Two sharp metamagnetic transitions are due to the phase transitions between the 'isotropic', (a) and (c), phases and the nematic state, (b). The inset shows the up-spin and down-spin Fermi surfaces in the nematic and 'isotropic' phases. Note that the down-spin Fermi surface volume gradually decreases as the Zeeman magnetic field increases, while the up-spin Fermi surface changes abruptly at the nematic-'isotropic' transitions. | cond | mat0408004-metamag.jpg |
0.492458 | 6d72ce9be97d4d598c27b1fce3024acd | The nematic order parameter, Δ_↑, as a function of Zeeman magnetic field, h, for F_2 N_0 =0.1 and μ = - 0.16 in unit of 2 t at T=0 | cond | mat0408004-order_new.jpg |
0.403037 | 21e0e9a94ce7430994e9c5a8afbfd99f | zt/U_0 dependence of ⟨ S_x⟩ for U_2/U_0=0.04 and μ/U_0=1.5 under a magnetic field gμ_B B=0.005. | cond | mat0408014-mu1.5mag.jpg |
0.486898 | 3232305d41e34f31b89c36a0da1ee2f9 | Phase diagram of the Bose-Hubbard model of spin-1 bosons for U_2/U_0=0.04. Here, z is the number of adjacent sites in the lattice. SF and MI indicate the superfluid and the Mott-insulating phases, respectively. The solid and dashed curves are obtained using the GW and the PMFA, respectively. The inset indicates the SF-MI phase boundary around the MI state with N=3 for U_2/U_0=0.001. | cond | mat0408014-phasediagram.jpg |
0.472576 | a3765b6e81fe40aa90cd700b473eec98 | Spin Rotator Chain. | cond | mat0408016-fig1d.jpg |
0.448143 | 7f61264c28124d1c85237ec42bc9ff41 | Dispersion law for hybridized spin fermions c_±. | cond | mat0408016-fig2b.jpg |
0.517618 | 45335a1811db4ccf9b87ae11febaabff | Schematic representation of the multilayer model. | cond | mat0408027-fig1.jpg |
0.483691 | 0f123b2212a442d698ef8cb5aadf1110 | Calculated spectra of the real (ε_bl1) and imaginary (ε_bl2) parts of the “intrabilayer dielectric function" ε_bl defined by Eq. (2) (a) and of the real part of the c-axis conductivity (b). | cond | mat0408027-fig2.jpg |
0.458781 | c30b2402ad6f4d83bcea96de02e70c7e | Top view of the slab after structural optimization, for the clean (left) and oxidized (right) Si(100)–p(2x2) surfaces. The x axis is perpendicular to the surface, and Si-Si surface dimers are oriented along the y axis. | cond | mat0408028-incze.figure.1.jpg |
0.428041 | 65880b90c71a4385bd10807d014a9da9 | Effects of the 0.5 ML oxidation on the RAS of Si(100)–p(2x2): clean surface (dotted line) versus oxidized one (full line). In the inset, the low-energy region is expanded. Besides the oxidized surface (full line) and the clean Si(100)–p(2x2) (dotted line), we also compare the RAS of the clean Si(100)–c(4x2) reconstruction (dashed line). A gaussian broadening of 75 meV has been used. | cond | mat0408028-incze.figure.4.jpg |
0.449007 | 70925682a2a14cecb111c222afe2ba67 | Side view of the slab, and the slicing used for the layer–by–layer spectral decomposition performed according to the method of Ref. <cit.>. Left: view axis perpendicular to surface dimers. Right: view axis parallel to surface dimers. Darker spheres are used for oxygen, grey ones represent silicon atoms. The oxidized region is fully enclosed in slice 5. | cond | mat0408028-incze.figure.5.jpg |
0.499304 | c0ed11f903c648168f26445b4478299f | Results of the layer-by-layer spectral decomposition, according to the slicing displayed in Fig. <ref>. The spectral feature at about 1.5 eV, recognized as an oxygen signature (see text), comes from the topmost slice only. In contrast, the main peaks at about 1 eV and 3 eV also include contributions from subsurface states (slice 3 and 4). | cond | mat0408028-incze.figure.6.jpg |
0.510928 | 006c6ccd101c42fcacd90db9a865547e | Contributions of selected slab states involved in the strongest optical transitions for Si(100)–p(2x2):O, below 1.7 eV. Three energy windows, corresponding to the main spectral features in the imaginary part of the half–slab polarizability tensor Im[α_ii^hs], are studied: A ([0.8,0.9] eV), B ([1.0,1.15] eV), and C ([1.40,1.75] eV). The full and dotted lines correspond to light polarization along the z and y directions, respectively. On the left, we represent schematically the slab states, below and above the Fermi level, and their originated transitions which carry the largest oscillator strength, and anisotropy, in regions A, B and C. | cond | mat0408028-incze.figure.7.jpg |
0.482307 | 4ef7fcc16ba5447abfad37bf02b5e83d | Contribution of the eight surface states singled out in Fig. <ref> to the Si(100)–p(2x2):O RAS. Dotted line: only 4+4 bands included in the summations of Eq. (<ref>); dashed line: full calculation with 108+250 bands. | cond | mat0408028-incze.figure.8.jpg |
0.466316 | 46d3f7cf461540ba8c7d5547fd6e2c51 | The blank squares and triangles depict the inner fluid of particles A and B respectively. The filled squares and triangles denote the discretized surface; fluid particles crossing that surface are reflected. The empty circles represent the outer fluid, while crosses (C) denote the shared boundary nodes. The filled circle E is not a real shared node, but belongs to both spheres (A and B), although it lies on different links. D is a single outer node between the two surfaces. If those surfaces move towards each other, high pressure occurs at D. | cond | mat0408029-fig1.jpg |
0.370148 | d839ef22126345999868e44e62a62a9b | The velocity of the “pseudo-wall-slip” versus shear rate for two different volume concentrations. The dependence of the velocity is linear. The slope of the line gives an effective width of the particle free region near the wall. The width is 1.16·10^-4 m and 9.23·10^-5 m for ϕ=0.026 and ϕ=0.053, respectively. Narrower particle free regions are caused by higher forces due to weight of particles being above the particle layer near the wall. | cond | mat0408029-fig12.jpg |
0.453606 | 2ff9f7035f294e13a9726b5df9631191 | Distribution of the particle velocitiy component parallel to the shear direction averaged over 5.55·10^6 time steps of the steady state. The peaks are separated by 2.5·10^-3 m/s, starting at 1.1·10^-3 m/s. Dividing this velocity difference (2.5·10^-3 m/s) by the shear rate (i.e. 10 s^-1) results in the particle diameter since the average width of the layer corresponds to one particle diameter. These layers move against each other with a relative velocity corresponding to the shear rate. | cond | mat0408029-fig14.jpg |
0.466939 | 1484dcb1faff46c29879270272ed4983 | Poisseuille-flow of a fluid with viscosity η = 1/9 and density ρ=1 under gravity g=10^-4 exerted on each lattice point, channel width L∈{8; 16; 32}. Gravity is set to g=5× 10^-5 for channel width L=16. The solid line represents the expected profile (Eq. (<ref>)). | cond | mat0408029-fig2.jpg |
0.462067 | 7a41abda76284964acf45ef0de204ca6 | Cut through a three-dimensional system after 5155 time steps. The link-bounce-back boundary conditions are implemented on the surface of the sphere and the walls. The particle Reynolds number is = 6 and g⃗ is the gravitational force. The movement of the interior fluid has already relaxed to solid body movement. | cond | mat0408029-fig3.jpg |
0.489977 | 28820943389f403a892347c365d7845e | Dependency of the correlation time τ_corr on the shear rate γ. All data points lie on a straight line, which indicates an exponential behaviour of τ_corr: τ_corr∝ e^-γ/γ_0, with γ_0=4.78 s^-1. | cond | mat0408029-fig8.jpg |
0.479839 | 7aac381577b14b0f953e1cf8a5e3346c | (a) Temporal evolution of the differential transmittance T/T. The magnitude of the spin-related slow component T/T_ mag is shown by horizontal lines. Inset shows the energy dependence of T/T. (b) Magnitude of the photo-induced demagnetization measured by time-resolved magneto-optical Kerr effect (right axis) and of T/T (left axis). | cond | mat0408030-fig1.jpg |
0.542841 | 5ec2cfe6590d42319d688770f00d1f13 | (a) Temperature dependence of the magnetization measured with a SQUID. Horizontal lines in the inset show the magnitude of the slow response of the differential transmittance T/T_ mag. (b) The solid line shows the magnetization change due to a 660 μ J/cm^2 heat injection (see text). Circles and squares show T/T_ mag at 190 and 250 meV probe energies respectively (the right vertical axis). The scaling factors for the two sets of data are different. | cond | mat0408030-fig2.jpg |
0.383749 | 2eb941dc049f407c9cf18c0f6e6febc3 | (a) Valence band configuration and midinfrared optical transition in Ga_1-xMn_xAs following the acceptor bound magnetic polaron (ABMP) model of carrier-induced ferromagnetism. (b) Schematic illustration of the spatial variation of hole energy and spin configurations below (left) and above (right) the Curie temperature according to the ABMP model. A deeper potential well (i.e., a larger binding energy) is obtained above the Curie temperature. | cond | mat0408030-fig3.jpg |
0.486996 | 7b96d6954b6349de9aeee71a06d74e07 | (a): Fitted (solid lines) and experimental (data points) curves of the linear absorption at 54, 110 and 290 K. (b) Simulated (solid line) and experimental (data points) variation of T_ mag/T as a function of the probe energy at 54 K. | cond | mat0408030-fig4.jpg |
0.413981 | 4301cfe863904d3a8c8c3673f982d515 | Average density ρ (x) (a) and current j (x) (b) for parameters α=0.8, β=0.8, Ω=0.3 and K=1. We use the same legend as in Fig. <ref>. The bulk density profile is given by the Langmuir density ρ_l=1/2 which corresponds also to the maximal current phase. Due to fluctuations, neglected in the mean-field approximation, the current profile obtained from the simulation exceeds the value 1/4 at each boundary. | cond | mat0408034-fig10b.jpg |
0.463673 | 8b122ac5381c4ea892e721da9728d33c | Cut of the phase diagram on the (α,β)–plane in the mean-field approximation for K=1 and different values of Ω: (a) Ω=0.3, (b) Ω=0.5, (c) Ω=1.0. The cases (a)-(c) correspond to the three different topologies of phase diagrams discussed in the main text. All lines represent continuous transitions between different regions in the (α, β, K = 1, Ω_D = const.) cut of the 4-dimensional parameter space. The line parallel to the anti-diagonal is defined through the relation α + β +Ω = 1. It represents the border line where the points x_α and x_β (i.e. the points where the left and right solutions ρ_α and ρ_β meet the Langmuir isotherm ρ_l=1/2) coincide, x_α=x_β. The phase boundaries of the LD-HD coexistence phase, x_w=0 and x_w=1, correspond to regions in which the domain wall is located at one of the system boundaries. These lines were computed by using the matching conditions for the currents: j_α(1) = β(1-β) and j_β(0) = α(1-α). In figure (a), we also emphasize the presence of the boundary layers at the left or the right end of the system. These are indicated with the letters "l" and "r", respectively. Such boundary layers remain present in the same regions as in figure (a) also for increasing Ω. | cond | mat0408034-fig11b.jpg |
0.475948 | 4ef8dedb6b6941eba956af321c1173e3 | The real branches W_0(ξ) and W_-1(ξ) of the Lambert W-function. | cond | mat0408034-fig12b.jpg |
0.439491 | f825e6e3041b42af997e8b265b6c30e9 | Mathematical solutions for (a) the left density ρ_α(x) and (b) the right density ρ_β(x) for K=3, Ω_D=0.1 and different values of the entry and exit rate α and β. The solutions which approach the Langmuir isotherm are those for α, β≤ 1/2 (thick lines). The solutions where the branching point coincides with the right boundary are indicated by α_c=0.038532... and β_c=1/2. | cond | mat0408034-fig13b.jpg |
0.457216 | 02feebace9e042eb9f183a45d3dd45bd | Average density ρ(x) (a)-(c) and corresponding current j(x) (b)-(d) for α, β≤1/2 in a parameter regime showing phase separation. We have chosen Ω_D=0.1, K=2 and (a)-(b) α=0.1, β=0.1 or (c)-(d) α=0.3, β=0.4. Solid lines correspond to the numerical solution of the mean-field theory with ε=10^-3. Monte-Carlo simulations are shown as solid wiggly line. The flat dashed line represents the Langmuir isotherm, ρ_l=K/(K+1). The other dashed lines represent the analytic solutions given by the branches of the Lambert W-functions matching the boundary conditions on the right and left end, respectively. For both cases (a) and (c), the solution matching the left boundary condition ρ_α is given by the branch of the Lambert W-function W_-1(-Y_α). For the solution matching the right boundary condition, ρ_β, one has to consider the branch W_0. For (a) the branch of W has the argument Y_β(x), while for (c) the argument is -Y_β(x) (see as illustration also Fig. <ref>). | cond | mat0408034-fig14b.jpg |
0.410541 | 3543034f7d154d1d9c8152fc33c5c11f | Average density ρ(x) (a) and current j(x) (b) for α=0.3, β=0.1, Ω_D=0.1 and K=2. We use the same legend as in Fig. <ref>. Except the left boundary layer, the bulk density profile is given by the Lambert W-function, ρ_β=W_0(Y_β(x)). | cond | mat0408034-fig15b.jpg |
0.398827 | b88b7fed54994391b07ad749397580cf | Average density ρ(x) (a) and current j(x) (b) for α=0.1, β=0.4, Ω_D=0.1 and K=2. We use the same legend as in Fig. <ref>. Except the right boundary layer, the bulk density profile is given by the Lambert W-function, ρ_α=W_-1(Y_α(x)). | cond | mat0408034-fig16b.jpg |
0.427489 | 2fcf2bef4ffc427b9ad820562c8b4254 | Average density ρ(x) (a) and current j(x) (b) for Ω_D=0.1, K=2, α=0.75 and β=0.75. We use the same legend as in Fig. <ref>. Except for the left and right boundary layers, the bulk profile obtained from the analytic mean-field result is given by the branch ρ_β(x)=W_0(-Y_β(x)) of the Lambert W-function computed for β=1/2. | cond | mat0408034-fig17b.jpg |
0.519223 | 0eee866403264516be379c47c9ea3ec3 | Cuts of the phase diagrams on the (α, β)–plane obtained by the exact solution of the stationary mean-field equation (<ref>) in the inviscid limit ε=0 for Ω_D=0.1 and (a) K=1.1, (b) K=3.0, (c) K=6.0. The two lines, corresponding to regions in which the domain wall is located at x_w=0 and x_w=1, are obtained by using the matching conditions for the currents: j_α(1)=β(1-β) and j_β(0)=α(1-α). In figure (a), we emphasize several features. With the letters "l" and "r" we indicate the presence of boundary layers in the average density profile, forming at the left or the right end of the system, respectively. In the lower left quadrant, the left and right boundary layers form whenever the domain wall exits the system on the left and right end side. At the phase boundary between the HD and M phases, for β=1/2, a boundary layer forms at the right end. Note that also in the M phase ρ_bulk > 1/2. The presence of boundary layers in the different phases of the (α, β)–plane is conserved upon variation of the binding constant K. The filled black circle represents the critical point C where the domain wall exhibits critical behavior; see Sect. <ref>. This critical point exits the plane for large values of K, accompanied with a topological change of the phase diagram. | cond | mat0408034-fig18c.jpg |
0.450801 | a909f779ad684afa86ffe9e56dac715a | Cuts of the phase diagrams as in Fig. <ref> for K=3 and (a) Ω_D=0.01, (b) Ω_D=0.05, (c) Ω_D=0.2. The white circle corresponds to a nodal point of the system N defined by the condition α=β=1-ρ_l=1/(K+1). Every line x_w=0 crosses this point for an increasing Ω_D. | cond | mat0408034-fig19c.jpg |
0.388863 | 5009b8e81dea472faf3e4fa44b49a621 | Illustration of the Totally Asymmetric Simple Exclusion Process with open boundaries. The entrance and exit rates at the left and right end of the one-dimensional lattice are given by α and β, respectively. | cond | mat0408034-fig1b.jpg |
0.406755 | 532ec9c26a544359baefb0044e3c9055 | (a)-(b) Domain wall position x_w and height Δ_w as a function of the entrance α for different values of the exit rate β at Ω_D=0.1 and K=3. At the critical point α=α_c and β=1/2 a domain wall forms at the right end of the system with an infinitesimal height Δ_w. The value of the "critical" entry rate is α_c=0.038532... and can be written explicitly by using the analytic solution in the mean-field approximation, see Eq. (<ref>). (c)-(d) Domain wall position x_w and height Δ_w as a function of the exit rate β for different values of the entrance rate α at Ω_D=0.1 and K=3. For α=α_c and β=1/2 a domain wall forms at the right end of the system with an infinitesimal height Δ_w. For exit rates β > 1/2, both domain wall position x_w and height Δ_w become independent of β. Changes in the exit rate only affect the size and shape of the boundary layer on the right end, but not the bulk density profile. | cond | mat0408034-fig20d.jpg |
0.433261 | 89c0fd027d1543f7815dca57650e9176 | Average density profiles computed analytically in the inviscid limit expressed in terms of Lambert W-function (dashed lines) and numerically for ε=10^-3 within a mean-field approximation (solid smooth line). Parameters are: α = α_c (see Eq. (<ref>) for the analytic expression), β = 0.5, Ω_D = 0.1 and different values of K. The profile is entirely given by the left solution ρ_α for the value of the entry rate α_c, defined by the condition ρ_α(x=1) = 1/2, and β = 1/2. Note that in this case, ρ_α matches simultaneously the left and the right boundary conditions. | cond | mat0408034-fig22b.jpg |
0.540988 | 577cf7d389964020b11092d89e6c8211 | Double decimal logarithmic plots of the critical behavior of the domain wall height, Δ_w, and position from the right end side, 1-x_w. We obtained the plot numerically with the program Maple, release 7, using the analytic mean-field solution in the vicinity of the critical point C and applying the matching condition over the left and right currents, j_α(x_w) = j_β(x_w). (a) As a function of α starting from the point C on the critical manifold with coordinates α_c = 0.038532..., β = 1/2, Ω_D = 0.1 and K = 3. (b) As a function of β starting from the point C on the critical manifold with coordinates α = 0.038532..., β_c = 1/2, Ω_D=0.1 and K=3. (c) As a function of K from the point C on the critical manifold with coordinates α=0.2, β = 1/2, Ω_D = 0.051443... and K_c = 3. (d) As a function of Ω_D from the point C on the critical manifold with coordinates α = 0.2, β = 1/2, Ω_D,c = 0.051443... and K=3. The value of Ω_D,c can be easily obtained from Eq. (<ref>) and the initial condition σ(0). Note the different scaling regime for the exit rate β. | cond | mat0408034-fig23d.jpg |
0.480081 | 9403c488e14b498893f48de76395f33a | Domain wall position x_w in logarithmic scale as a function of Ω_D at α=0.2, Ω_D=.051443... and K=3 and different values of β. If β > α the domain wall builds up from the right boundary, while for β<α from the left boundary. For α = β the domain wall approaches the position x_m which is independent of the decreasing detachment rate Ω_D. At large Ω_D the domain wall position x_w always moves to the left boundary as 1/Ω_D. | cond | mat0408034-fig24b.jpg |
0.456094 | 513d535f12ff41609c06bc1061e37b6f | Domain wall slope S_w estimated from Monte Carlo simulations as a function of the system size N. Simulations where performed for α=0.2, β=0.6, K=3 and Ω_D=0.1. | cond | mat0408034-fig25b.jpg |
0.50509 | 7cf6e5ea2de4429cb849953fa0248c10 | Illustration of Langmuir kinetics. ω_A and ω_D denote the local attachment and detachment rates. | cond | mat0408034-fig2b.jpg |
0.48093 | de5c75093a404a94a62e7d19f7b9266b | Schematic drawing of the totally asymmetric simple exclusion process with bulk attachment and detachment <cit.>. The entrance and exit rates at the left and right end of the one-dimensional lattice are given by α and β, respectively; ω_A and ω_D denote the local attachment and detachment rates. | cond | mat0408034-fig3b.jpg |
0.427095 | e34d824e424349aba6fe6d0a63b2cfa9 | Illustration of the network architecture corresponding to the totally asymmetric simple exclusion process (TASEP) and Langmuir kinetics (LK). | cond | mat0408034-fig4b.jpg |
0.441762 | b4843c89a7cd4504a4b84bf236628b06 | Average density ρ(x) (a) and current j(x) (b) for parameters α=0.4, β=0.4, Ω=0.3 and K=1. In this parameter range one observes a 3-phase coexistence: a maximal current phase is intervening between a low and high density phase. The profiles are computed analytically in the inviscid limit (dashed lines) and numerically for ε=10^-3 within a mean-field approximation (solid smooth line), and from Monte-Carlo simulations (solid wiggly line). Note that, within the resolution of the figures, the Monte-Carlo results and the numerical mean-field results can not be distinguished. The analytic density profile is shown for the solutions respecting the left and right boundaries conditions, ρ_α and ρ_β; we also show the Langmuir isotherm ρ_l=1/2. | cond | mat0408034-fig5b.jpg |
0.406397 | 1347b0fb263f4c7ca839d1cb6dae6c11 | Average density ρ (x) (a) and current j (x) (b) for parameters α=0.4, β=0.1, Ω=0.3 and K=1. We use the same legend as in Fig. <ref>. The bulk profile is almost completely described by the solution ρ_β matching only the right boundary condition. At the left end, the bulk density does not match the boundary condition. As a result, a boundary layer appears. Only there does one find a noticeable difference between the profiles of the Monte-Carlo simulation, the numerical computation at finite ε and the analytic profile for vanishing ε. | cond | mat0408034-fig6b.jpg |
0.498281 | 32d1eebdc820437bb46f37c7ce9cc1d9 | Average density ρ (x) (a) and current j (x) (b) for parameters α=0.2, β=0.1, Ω=0.3 and K=1. We use the same legend as in Fig. <ref>. Only in proximity to the domain wall the results from the mean-field approximation show deviations from the density profile obtained by Monte-Carlo simulation. | cond | mat0408034-fig7b.jpg |
0.468868 | 1c7269bb489844a09c3941a4c8296f42 | (a)-(b) Average density ρ (x) and current j(x) for α=0.8, β=0.35, Ω=0.3 and K=1. We use the same legend as in Fig. <ref>. Except for the left boundary layer, in fig. (a) the analytic solution is described by the Langmuir density ρ_l=1/2 and the density ρ_β matching the right boundary condition. (c)-(d) Average density ρ (x) and current j(x) for α=0.35, β=0.8 and the same Ω and K as before. Note that the curves map to those of (a)-(b) by particle-hole symmetry. | cond | mat0408034-fig8c.jpg |
0.425261 | 9a78d6b8803b459e82775e4545910ae0 | Brush phase diagram, strong systems with salt. The < region (not shown) maps onto =. Arrows indicate how boundaries evolve with increasing salt concentration. (a) L>ξ>, (b) ξ<. Salt-dominated regions (shaded) are blown up at right. QB (Quasineutral brush) and QIC (quasineutral independent) are subdivided into bands α,β,γ. Since <, γ does not exist for ξ> (see text). | cond | mat0408036-evolution.jpg |
0.481797 | f80e6870d34d4ddca1e5d284407c95c2 | Brush phase diagrams for no added salt (grafting density b^-2, backbone charge separation l, logarithmic axes). (a) Weak and (b) strong systems. OB: osmotic brush; PB: Pincus brush; IC: independent chains; MC: Manning condensation; PMC: partial MC; FR: fully relaxed chains. Saturated phases are indicated by prefix “Sat”. | cond | mat0408036-nosalt.jpg |
0.423304 | 31e9899f92a14b6881c56ee1d4d87484 | The quasineutral brush (QB) behaves as a neutral brush with effective hard core monomer size depending on polymer charge density and salt concentration. There are 3 regimes, giving 3 bands in the phase diagram, fig. <ref>. Band α: length saturation, ≈ξ. Band β: string of electrostatic blobs, ≈ξ. Band γ: overscreening (ξ<), ≈^3/ξ^2. | cond | mat0408036-qb.jpg |
0.460658 | 61f584c527e54fc3aa36e6eb983b89cf | Time dependence of the external magnetic field during the field overshoot. The time origin corresponds to the nominal field stop. A triangular overshoot (dashed-dotted line) and an overshoot given by the function Δ H_ov = H_ov (t/t_ovm)^2 exp (2(1-t/t_ovm)) (dotted line) are shown. In the inset, the time derivatives of the two overshoot functions are shown. | cond | mat0408045-fig1.jpg |
0.432193 | 305eb27ce55c45b28c489c5926a10e30 | Magnetic field ramp with a sweep rate of 3.3 mT/s in the time window where a field overshoot occurs, as measured by a hall-probe (square and line), and H_a(t) employed in our computation (dotted line). | cond | mat0408045-fig10.jpg |
0.484354 | 9c574c30155747588b7663c956c193ee | Magnetic relaxations measured at different temperatures for a magnetic field H_0 = 1 T (points) and computed curves (continuous lines). | cond | mat0408045-fig11.jpg |
0.560481 | a1b3fb993af74f59b4231aeb1ccbde6a | Field profiles computed for the relaxation at 45 K. The profile in the direction of the arrow are computed at t=100, 1000, 5000 s. | cond | mat0408045-fig12.jpg |
0.433923 | 27c87b93e0bc40dd9d07185e8da31e12 | Magnetic relaxation curves computed for different magnetic field ramps in a thick strip. | cond | mat0408045-fig2.jpg |
0.448023 | 69c140af45b846b0830aacd2fc045772 | Magnetic relaxation curves computed for different magnetic field ramps in a slab. The dotted line is the relaxation given by the analytical formula reported in the text. | cond | mat0408045-fig3.jpg |
0.420219 | b4e69063ab4a441281d4f6160cfe425d | Time evolution of the magnetic field profiles in a slab; the initial field profile is achieved without field overshoot. | cond | mat0408045-fig4.jpg |
0.461869 | f086fbab5fdc44bc88b3d1fdd0e98553 | Time evolution of the magnetic field profiles in a slab, when a field overshoot occurs. The dotted lines represent the flux profiles which fully resemble the ideal profiles. The dashed lines represent the profile during field ramp rate reversing. Continuous lines show the field profiles in the time windows where the magnetization is nearly constant. In the inset a detail of the profile close to the slab surface is shown. | cond | mat0408045-fig5.jpg |
0.428776 | 89c529ace6ad490ebe6685c68ea6cf68 | Magnetic relaxation curves computed from different magnetic field ramps in a thick strip. | cond | mat0408045-fig8.jpg |
0.477865 | 056d2c0d4b4f4063aa93e6e221fb4c42 | Thermodynamic critical magnetic field H_c(T) normalized by H_c(T=0)_BCS. The bold dashed curve indicates data from ref.<cit.> and the bold solid curve indicates those obtained using the present two-band anisotropic model. The thin solid curve indicates the BCS results. The results for the σ-band (ab-plane anisotropic) and the π-band (prolate form) are also shown. | cond | mat0408054-Hc.jpg |
0.45044 | 0f84b1b3da41457b982c07e886c9280c | Reflectance spectra R(ω) of NbN_1-xC_x thin films deposited on MgO substrates at T=4.3 K (solid line) and T=20 K (dotted line). | cond | mat0408054-nbnref.jpg |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.