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
|
---|---|---|---|---|
0.457103 |
37e3cabe5ea74d2b9718dc1529f37da9
|
Low-energy structures of the 2a√(6.5)× 2a, 406-atom surface unit cells (rectangles) for Si(105). The surface energies γ are indicated next to the corresponding model labels. Atoms are rainbow-colored according to their height, as described in Fig. <ref>.
|
cond
|
mat0408228-plateaus406.jpg
|
0.513284 |
6fffc2a2ee2e4569b012c8e087b26931
|
Left: 2D billiard with ℓ=1. Right: Corresponding Aharonov-Bohm cylinder.
|
cond
|
mat0408240-rng_fig1.jpg
|
0.424154 |
7f18b32ac0ac4871aed96d461b15c19f
|
The probability distribution |⟨ν,μ |n ⟩|^2 for (a) The n=2423 eigenstate of the smooth (ℓ=1) billiard; (b) The n=1000 eigenstate of the rough (ℓ=100) billiard. Note that this is essentially the (p_x,p_y) momentum distribution. The state in panel (a), unlike the state in panel (b), is a typical semiclassical state. Namely it is well concentrated on the energy shell.
|
cond
|
mat0408240-rng_fig4.jpg
|
0.464138 |
21a6f1845eda48978fbfec6773bf6b73
|
(a) Equations of state for the columnar phase. The solid curve is the prediction from the present cell model, the dashed one is the traditional cell pressure, given by Eq. (<ref>). The symbols correspond to the simulation data from Zhang et al. <cit.> for D/L=10 (crosses) and D/L=20 (triangles). The inset shows the pressure in the dilute regime near the columnar-nematic (“melting") transition at ϕ^∗≈ 0.48. (b) Dimensionless chemical potential βμ̃=βμ - ln[ v_0/𝒱ϕ_cp] from simulations (dotted curves) and cell theory (solid curves). The prediction from the traditional cell model, indicated by the dashed curve, is independent of D/L.
|
cond
|
mat0408245-extrafig1.jpg
|
0.441589 |
5ad96770b62c48c0a9824986617979b9
|
Normalized inter- and intracolumnar spacings, Δ_c/D and Δ_n/L respectively, as a function of ϕ^∗. Solid lines are theoretical predictions, the symbols follow from simulations for D/L=10 (crosses) and D/L=20 (triangles). The dotted curve follows from the traditional cell model and denotes both spacings. Inset: Average polar angle relative to the internal angle (θ_int∼ L/D) plotted versus ϕ^∗.
|
cond
|
mat0408245-extrafig2.jpg
|
0.458965 |
f9a96c7575c3472592aef01e57c60c25
|
The performance of the `2nd neighbor' strategy. As before, the plot shows the changes in the ratio of the largest cluster in the network. In the short run, there is a steep decrease in the size of the largest component, but after that, the largest component starts to grow and reaches a quite high level. The 2nd neighbor strategy performs well in the long run. Also note the quite big fluctuations, even after the curve somewhat levels off.
|
cond
|
mat0408248-nei_giant_time.jpg
|
0.417707 |
76bd492d32084fe9a2d6c55e0a2aa089
|
The performance of the 2nd neighbor strategy in the function of the average node degree in the network. There might be a phase transition in the network, where an infinite giant component appears in infinite networks around average node degree 2 with probability 1.
|
cond
|
mat0408248-nei_pht.jpg
|
0.468407 |
6dcea6f6d09a4bf699e70dd0e2d5fd9c
|
The degree distributions of a random graph (blue line, circles) and a random graph after attacks and random rewiring repair (red line, squares). The plot shows that the random rewiring strategy does not keep a random graph random. This is because the attacks remove preferentially the nodes with high degree.
|
cond
|
mat0408248-rnd-dds.jpg
|
0.466231 |
732462925b674971845a2d415238008c
|
Phase transitions in random graphs (green line, squares) and attacked random graphs (read line, circles). Further calculations are needed to prove the existence of the transition in the attacked graph. The transition threshold for the attacked network seems to be higher.
|
cond
|
mat0408248-rnd_cluster_pht.jpg
|
0.46834 |
7910fdaf8934430a85ccd7320a85b1ae
|
The performance of the random (blue line, circles) and greedy (red line, squares) strategies starting with a random graph. The plot shows the ratio of the number of nodes in the largest connected component to the network size. If the random rewiring is applied, the network reserves the size of the largest component, even if the structure of the initial random graph changes. The greedy strategy always ends up with completely disconnected nodes after sufficiently many attacks.
|
cond
|
mat0408248-rnd_greedy_time.jpg
|
0.502982 |
2a1ea5832b614d6282a827401ce2c7fd
|
Charge and spin configurations of a t-J model in a pinned checkerboard configuration. The results show the expectation value of <S_z> and <n_h> for a 10x8 t-J cluster with cylindrical boundary conditions: open in x, periodic in y. Here there are 8 holes and J/t = 0.35. Initially, a local pinning potential of -0.5 t was placed on the 16 sites making up 4 the placquettes visible in the left panel. Nine sweeps were performed, keeping up to m=1000 states, with the result shown in the left panel. Subsequently, the pinning potential was turned off. In the right panel, we show the result after three more sweeps were performed, reaching m=1500 states.
|
cond
|
mat0408249-fig1.jpg
|
0.470984 |
cf5304c43d9c44f2832250b9a1444481
|
Charge and spin configurations of a t-J cluster with two sites pinned with a local potential of -0.5t. In each panel, the sites pinned have the large circles. The lower plot shows the total energies of the systems as a function of the number of states kept per block as DMRG sweeps were performed on the systems.
|
cond
|
mat0408249-fig2.jpg
|
0.548587 |
d2355f3cc5a54c1099577fc52d6ca8ca
|
Charge and spin configurations with different values of the exchange coupling linking vertical and horizontal bonds. In both cases J_x = 0.35. In the left panel, J_y = 0.37, and in the right, J_y = 0.38.
|
cond
|
mat0408249-fig3.jpg
|
0.519158 |
3703ba4ec2de4805bb3c43dd7b3cf199
|
Charge and spin configurations with a local field in a pattern transverse to the stripes applied. In the simulation shown in the left (right) panel, potentials of -0.05t (-0.1t) were applied to four of the horizontal rows of sites: rows 2,3,6, and 7.
|
cond
|
mat0408249-fig4.jpg
|
0.450843 |
b0bbabea91db4e7c8b5e6dab2224f485
|
(A) T dependence of the in-plane resistivity ρ_ ab (T) measured on cooling for Ca_2-xSr_xRuO_4 . Vertical broken lines are guides to the eye. (B) Phase diagram for the region 0 ≤ x ≤ 0.2 with three different phases: paramagnetic metal (blue), paramagnetic insulator (red), and antiferromagnetic insulator (green). T_ M-I and T_ AF are the M-I and AF transition temperatures determined on cooling. Thick and thin lines indicate the first and second order transitions.
|
cond
|
mat0408250-Figure1.jpg
|
0.497066 |
1151e459db2742fdafed04ee36a94c5c
|
T dependence of the in-plane susceptibility for Ca_2-xSr_xRuO_4 with x = 0.06, 0.09, and 0.15. All curves were measured in a field-cooled sequence. The solid and open symbols indicate the results measured on cooling and heating, respectively. Inset: (I) thermal hystereses observed in the paramagnetic phase for x = 0.06 and 0.09. (II) C_P/T vs. T^2 for x = 0 and 0.15. The solid lines represent linear fit.
|
cond
|
mat0408250-Figure2.jpg
|
0.501957 |
0189da2ad3274109b9c0abfbea1f7d70
|
(A) lnρ vs. 1/T^1/2 and 1/T^1/3 for the insulating phase of Ca_2-xSr_xRuO_4. The results for x ≤ 0.09 are plotted against the lower horizontal axis 1/T^1/2, while these for x = 0.15 are plotted against both 1/T^1/2 and the upper axis 1/T^1/3. Straight lines indicate a fitting result to Eq. (<ref>). (B) Corresponding plot for theoretical predictions. The results for W = 0.02 and 0.08 eV are plotted against the lower axis 1/T^1/2, while these for W = 0.16 eV are plotted against both 1/T^1/2 and the upper axis 1/T^1/3. Inset: (I) T_0 obtained by fitting of ρ(T) to Eq. (2) with α = 1/2 for Ca_2-xSr_xRuO_4 (solid circle) as a function of x (lower axis), and for theory (open circle) as a function of W (upper axis). (II) Model band structure used in our calculation. As the disorder increases (W : 0.02 eV → 0.16 eV), the band-tails gradually fill-in the Mott gap, producing a soft gap similar to that predicted by Efros and Shklovskii, but on a much larger energy scale.
|
cond
|
mat0408250-Figure3.jpg
|
0.421588 |
b288db8aaf5045b9abcda825096f8315
|
Standing wave granular patterns in laboratory experiments and molecular dynamics simulations for the same number of particles, 60,000, which fill a 100σ× 100σ container to a depth of 5.4 layers: (a) squares, (b) stripes, (c) and (d) alternating phases of hexagons, (e) flat layers, (f) squares, (g) stripes, and (h) hexagons. Patterns (a)-(e) oscillate at f/2, (f)-(h) at f/4. The experiment used lead particles with σ=0.55 mm. From <cit.>.
|
cond
|
mat0408252-MDexpt.jpg
|
0.443625 |
1c25e445fdb74e7fbbb75f0dfdc9bd21
|
(a) Electromechanical shaker system for studying pattern formation in vertically oscillating layers. (b) Spatial patterns are illuminated from the side by light incident at low angles. When viewed from above, high regions are bright and low regions are dark. From <cit.>.
|
cond
|
mat0408252-apparatus_new.jpg
|
0.512437 |
6b3c42ae4e5d4ad0850aa4d9f898ed94
|
Trajectory of a completely inelastic ball on an oscillating plate. This is a model for the motion of the center of mass of a granular layer. The sinusoidal curve is the trajectory of the plate. The ball leaves the plate when the acceleration of the plate becomes -g, that is, when the dot-dashed line intersects the trajectory of the ball. If the ball collides with the plate above the dot-dashed line, as in (b) and (d), it leaves the plate immediately. From <cit.>.
|
cond
|
mat0408252-ball.jpg
|
0.423353 |
d19c036018cb48469fc3a340a04ebd5d
|
The horizontal velocity field measured for the expansion fan that formed when the supersonic granular flow reached the bottom of the wedge. The solid lines indicate selected streamlines. The total height of the region shown is 55σ. The white region below the wedge had too few particles for the velocity to be determined. (From <cit.>)
|
cond
|
mat0408252-expansion.jpg
|
0.456902 |
994e423dc0404698b15bae968a4f9691
|
Forced granular materials produce qualitatively similar patterns as forced fluids: (a) stripe pattern formed by a vertically oscillated granular layer <cit.>, (b) stripe pattern formed by a vertically oscillated layer of water <cit.>, (c) stripe pattern formed in thermal convection of a fluid (CO_2) <cit.>.
|
cond
|
mat0408252-faraday_waves.jpg
|
0.471284 |
a131553b7e2143bead82496e3bddca7c
|
Instabilities of patterns found in oscillating granular layers and Rayleigh-Bénard convection in a fluid. Cross roll instability in stripes: (A) Vibrated granular layer <cit.> and (a) Rayleigh-Bénard convection <cit.>. Skew varicose instability in stripes: (B) Granular layer <cit.> and (b) Rayleigh-Bénard convection <cit.>. Spiral defect chaos in: (C) vibrated granular layer <cit.> and (c) Rayleigh-Bénard convection <cit.>.
|
cond
|
mat0408252-fig2_rbgrab2.jpg
|
0.404349 |
75a5227d69674184b0c2bd079372121f
|
Left: close up snapshot of a granular lattice (Γ=2.9, f=25 Hz, h=4σ). Right: time evolution of the peaks in the boxes A and B in the left-hand image. The peaks oscillate out of phase with a frequency about twenty times smaller than f=1/τ. (from <cit.>).
|
cond
|
mat0408252-latticeosc_new.jpg
|
0.478061 |
f0176c137a714106858d642456f462a6
|
Defect creation and melting of a square granular pattern in a vertically oscillating granular layer: (a) experiment and (b) molecular dynamics simulation. The insets show Fourier transforms of the spatial patterns. In the experiment the plate oscillation frequency (f=32 Hz) was modulated at the natural oscillation frequency of the lattice (2 Hz), and at t=τ graphite powder was added to the layer of bronze spheres. By t=56τ defects had formed, and by t=175τ the lattice had melted. In the MD simulation the friction coefficient was reduced from μ=0.5 to zero at t=τ; by t=22τ defects had formed, and by t=100τ the lattice had melted. (Γ=2.9.) From <cit.>.
|
cond
|
mat0408252-melting_new.jpg
|
0.501793 |
d25d03aacf4b40f9be542e4de3084cbc
|
(a) Dimer formed by two bound oscillons of opposite phase; one period of oscillation of the container later the white peak will have become a crater and the black will have become a peak. (b) Tetramer formed of four oscillons. (c) Polymer chain of five oscillons. (d) A square lattice grows by nucleating oscillons. Individual bronze spheres (σ=0.165 mm) are discernible in (a)-(c). From <cit.>
|
cond
|
mat0408252-os_figure3_new.jpg
|
0.425309 |
c3ca1969269243cab0293aa642cfb233
|
(a) Snapshot of a container with two oscillons (viewed from above); one is a peak (upper left) and the is other a crater (near the center). (b) and (d), Close-ups of an oscillon crater viewed from the top and from the side, respectively. (c) and (e) Close-ups of an oscillon peak viewed from the top and from the side, respectively. Individual bronze spheres (σ=0.165 mm) are discernible in (b)-(e). (f=25 Hz, Γ=2.45, layer depth h=17σ.) From <cit.>.
|
cond
|
mat0408252-oscillons_new.jpg
|
0.461599 |
3405c0e20c254e3b8ce2ad3eccc6d47e
|
Patterns in oscillating granular layers: (a) stripes, (b) squares, and (c) hexagons (with different phases on the left and right). The patterns oscillate at f/2, where f is the container oscillation frequency. Here Γ, f, and h/σ (ratio of the depth of the layer at rest to the diameter of the particles) are given by (a) 3.3, 67 Hz, 7, (b) 2.9, 25 Hz, 4, and (c) and (d) 4.0, 67 Hz, 7. In (a) and (c) the diameter of the container is 770σ and in (b) the container is square, 1100σ× 1100σ. The particles are bronze spheres 0.165 mm diameter. (a) and (c) are from <cit.> and (b) is from <cit.>.
|
cond
|
mat0408252-patterns_new.jpg
|
0.435747 |
5221823e238f40e5ad9123f69716bd6d
|
Phase diagram for granular patterns observed in a vertically oscillated container, as a function of the dimensionless acceleration Γ and dimensionless frequency f^* = f√(h/g). The transitions from a flat layer to squares are hysteretic: solid lines denote the transition for increasing Γ while dotted lines denote decreasing Γ. (Bronze spheres, σ=0.165 mm; layer depth, 5.0σ; container diameter, 770σ.) From <cit.>.
|
cond
|
mat0408252-phasediagram.jpg
|
0.427524 |
c7304d39361b47c8be06f85300d65afa
|
Dimensionless temperature T/gσ and volume fraction ν as functions of the dimensionless height z/σ at four times ft in the oscillation cycle. For each time, a snap shot from the MD simulation is shown in the left column, with individual particles color coded according to temperature: high T in red, low T in blue, and the bottom plate of the container shaded solid gray. The right column shows horizontally averaged ν (blue) and T/gσ (red) for the same four times. The plate is shown as a horizontal black solid line, results from MD simulation are shown as dots, and continuum results are solid lines (From <cit.>).
|
cond
|
mat0408252-prop_shock.jpg
|
0.398873 |
f62653a7ae654a068323bd6817bc87fe
|
Location of the shock (solid line for continuum, squares for MD) and the center of mass of the layer (dashed line for continuum, circles for MD) as a function of time ft during one cycle of the plate (thick solid line) for particles with e_0=0.90. The plot is shaded according to the volume fraction from the continuum simulation, so that high volume fraction is dark and low volume fraction is light. The “top” and the “bottom” of the layer from MD (when the volume fraction drops to less than 4% of that for random close packed particles) are shown as +'s and X's respectively. The material below the shock is compressed as compared to the region above the shock, as can be seen from the shading. From( <cit.>).
|
cond
|
mat0408252-shock_location.jpg
|
0.432165 |
aec27c518e8c42f79899f047de7fcdcd
|
Average dimensionless shock speed, v_shock/√(gσ), in the reference frame of the plate. v_shock is calculated as the average speed of the shock from when the shock is formed until it leaves the layer. (From <cit.>)
|
cond
|
mat0408252-shock_speed.jpg
|
0.553864 |
4c344e56d64c48fdaaa0bf6eb796c572
|
Shock profiles for granular flow past a wedge measured in an experiment (circles) are compared with results from molecular dynamics (solid lines): (a) volume fraction, (b) horizontal component of the velocity, and (c) temperature. The profiles are taken along the dashed line in fig. <ref>. (From <cit.>)
|
cond
|
mat0408252-wedge_plots_friction_new.jpg
|
0.454352 |
be0f143e2f304dc49de4eda44b6de713
|
Horizontal component of the velocity field of a granular flow incident downward on a wedge, determined by three methods: (a) experiment, (b) MD simulations, and (c) integration of inelastic continuum equations. Each picture shows a region 130σ by 104σ. The solid lines with arrows denote streamlines. Quantitative comparisons along the dashed line in (a) are shown in figs. <ref> and <ref>. (From <cit.>).
|
cond
|
mat0408252-wedge_shock.jpg
|
0.408819 |
6c18ebe0f6cd4b0384a312431e9462f4
|
Comparison of shock profiles for granular flow past a wedge obtained from molecular dynamics (solid lines) and inelastic continuum equations (dotted line), assuming no friction. (a) Volume fraction, (b) horizontal velocity profile, and (c) temperature along the dashed line in fig. <ref>(a). Experimental measurements (open circles) show similar qualitative behavior but disagree quantitatively. The difference between the simulations and the experiment is due to wall friction. (From <cit.>)
|
cond
|
mat0408252-wp_no_friction_new3.jpg
|
0.468535 |
6188b10452854f91a491c473be32b604
|
Energy structure and the scattering lengths of ^6Li_2 in the (1,3) and (1,2) channels (dotted lines). All energies are referenced to the (1,2) scattering threshold. In these two channels, Feshbach couplings induce the formation of molecules (solid lines) below 69.1mT and 83.5mT, respectively. Arrows show the bound-bound and bound-free transitions based on molecules in the (1,2) channel. The lower figure shows the scattering lengths in the two channels.
|
cond
|
mat0408254-energy_structure.jpg
|
0.428357 |
2df701fd5d8540528cadcc5e77e3155d
|
RF spectra of ^6Li_2 molecule at different magnetic field values. Theoretical curves are based on Eq. (<ref>) and Eq. (<ref>) with the binding energies and scattering lengths obtained from the multichannel calculation. To shown both the bound-free and bound-bound spectra, an integration bandwidth of 1kHz is assumed. RF-induced loss on the molecular population at 76.4mT (solid squares), 72.0mT (open squares), 69.5mT (solid circles), 68.4mT (open circles), 67.6mT (solid diamonds) and 66.1mT (open diamonds) are re-scaled. Experiment data are from the Grimm's group in Innsbruck.
|
cond
|
mat0408254-exp.jpg
|
0.534623 |
488d07f110a54eafb382bbbe0bb3b582
|
Fano profile near a Feshbach resonance in the outgoing channel A'. The peak Franck-Condon factors with E_b=1 are calculated for different a/a' based on Eq. (<ref>). Near the Feshbach resonance, a/a' approaches zero.
|
cond
|
mat0408254-fano.jpg
|
0.487634 |
0a03175141824521beda10e1ffba0ebc
|
Comparison of the Franck-Condon factors from numerical calculation and from theory. Numerical calculation at 72.0mT (open squares) and 68.0mT (open circles) are plotted together with the formula Eq. (<ref>) (solid line at 72.0mT and dotted line at 68.0mT). The parameters are E_b=k_B× 16.2μK and a'=25173a_0 at 68.0mT; E_b=k_b× 6.2μK and a'=-7866a_0 at 72.0mT.
|
cond
|
mat0408254-fcf.jpg
|
0.514067 |
160920ebc1b9473f929ec641fd381184
|
Transition from a bound-free transition to a bound-bound transition with E_b=1. For the bound-free transition, the Franck-Condon factor is plotted as a function of K/E_b. Curves of different a'/a are offset and shown in the order of that near a Feshbach resonance in the A' channel. (dotted line for the resonance condition) For a'>0, the locations of the bound-bound transitions are indicated by the sharp peaks. Notably, bound-free transition vanishes at a'=a.
|
cond
|
mat0408254-lineshape.jpg
|
0.518406 |
7dc1ae5090ff4495b4aa636e266537ea
|
Comparison of the scattering phase shifts from numerical calculation and from the effective range expansion. The numerical calculation (open circles) is shown together with the predictions from Eq. (<ref>) with (solid line) and without (dashed line) the effective range correction. The scattering parameters are based on ^6Li in the (1,3) channel near 72.0mT with a_13=-7866a_0.
|
cond
|
mat0408254-phase.jpg
|
0.541003 |
73f96eb7ae7b40f1abc7234f7404d54f
|
Inflation rule of the octagonal QC (tiling): The inflated tiles are sided by thicker lines. The rhombic tile after the inflation is symmetrically decorated with the original tiles but the square one is asymmetrically. The square has a polarity along a diagonal and the polarity is determined by the presence of twinned rhombi sharing sides with the square.
|
cond
|
mat0408255-Fig1.jpg
|
0.456022 |
5ce653a62550438ab01aa56ebb3c4658
|
A self-similar octagonal tiling in 2D: It is produced by an inflation rule for the three kinds of tiles. The left-bottom corner is the center of the exact octagonal symmetry. This center is the origin of E_2, so that it is an even site. A part of the inflated tiling is shown at the left-bottom, while a part of the doubly inflated one is at the right-bottom.
|
cond
|
mat0408255-Fig2.jpg
|
0.539316 |
d91e2cde15e64957a8d6d3e1028c6767
|
A 1/8 sector of the structure factor of the octagonal SQC: Reflections with lower intensities than 0.001 in unit of the maximal value are ignored. The blue spots show main reflections, while those in green and red show the first-order superlattice reflections and the second-order ones, respectively. The area of a spot is proportional to the relevant intensity.
|
cond
|
mat0408255-Fig3.jpg
|
0.571628 |
2467418538c945aea680ed82bb11c6c0
|
The configuration space of the system has the shape of a dumbbell with two regions X, Y which are not directly connected by the Hamiltonian. Consequently, a trial wavefunction which has value -Ψ_0 on Y and Ψ_0 on X has vanishing energy in the thermodynamic limit.
|
cond
|
mat0408257-dumbbell2.jpg
|
0.442153 |
e0d020e478054445858cc1bc076b1ed7
|
The catenoid surface defined by r(z)=acoshz with a=1. The value of a corresponds to the inner radius of the pore in unit of the characteristic pore depth along z.
|
cond
|
mat0408258-fig1.jpg
|
0.546643 |
e7e4955fb26941e287a6d595ee43ff4a
|
Effective potential energy curves W_m(x) for catenoid when there is no magnetic field. These four curves corresponds to m=0 (bottom), m=± 1, m=± 2, and m=± 3 (top), respectively.
|
cond
|
mat0408258-fig2.jpg
|
0.415979 |
4c08d52123cb4293b31e66f8ed526b44
|
The dependence of the eigenenergies of the bound states on the Aharonov-Bohm flux through the pore of the catenoid, for (a) a=0.25, (b) a=1.0, and (c) a=2.5. Three congruent curves (shifted horizontally) correspond to m=-1, 0, and 1. For a=0.25, there exists the second bound state for each m and is shown by a broken line.
|
cond
|
mat0408258-fig3.jpg
|
0.480914 |
e224299596f54174b03bdc5608ec9569
|
A sinusoidal surface defined by r(z)=a+bcos z with a=1.5 and b=0.5.
|
cond
|
mat0408258-fig4.jpg
|
0.471066 |
a855f29d3c054adc83b6a4abd7c2bbf5
|
Effective potential energy curves W_m(x) for the sinusoidal surface (a=1.5, b=1.0) when there is no magnetic field. These four curves corresponds to m=0 (bottom), m=± 1, m=± 2, and m=± 3 (top), respectively.
|
cond
|
mat0408258-fig5.jpg
|
0.457317 |
fda93fb242c04d7392e00ebff3ad6307
|
(left) The energy band structures for a sinusoidal surface with a=1.5 and b=1.0 without magnetic field, in which the values of m corresponding to each band is shown in parenthesis. Note that for fixed m, energy gaps are the most prominent in a lowest part of the spectrum, and it becomes dominant for larger m. (right) Charge density distributions along the sinusoidal surface with three different Fermi energies E_F=3.0, 3.5, and 4.0, corresponding to the number of electrons per period N_e=35.39, 41.49, and 51.61, respectively. The brighter color corresponds to high concentration of the charge density.
|
cond
|
mat0408258-fig6.jpg
|
0.495137 |
66efb260c1bc4ce7b8896cb0b23eaba0
|
The charge density distribution calculated for E_F=3.0 with two different values of magnetic flux, (a) Φ=0.3Φ_0 and (b) Φ=0.5Φ_0.
|
cond
|
mat0408258-fig7.jpg
|
0.403792 |
9918a92eeae6476aa5af915a0306863a
|
Definitions. In the KJMA model, a hole is the liquid domain between the growing solid domains (island). The island-to-island is defined as the distance between the centers of two adjacent islands.
|
cond
|
mat0408260-fig1.jpg
|
0.571371 |
daf132466faf40e6877fbaef4825ef17
|
Kolmogorov's method. (a) Spacetime diagram. In the small square box, the probability of nucleation is I_0 Δ x Δ t, where I_0 is the nucleation rate. In order for the point X to remain uncovered by islands, there should be no nucleation in the shaded triangle in spacetime. (b) Kinetic curve for constant nucleation rate I_0: f(t) = 1- exp(-I_0 v t^2).
|
cond
|
mat0408260-fig2.jpg
|
0.414841 |
d0db0d8fd92244189d9b8c3b63e24360
|
Spacetime diagram. The hole-size distribution ρ_h(x,t) is proportional to the probability p_0(x,t) for no nucleation event occurs in the shaded parallelogram ABCD (see text).
|
cond
|
mat0408260-fig3.jpg
|
0.365437 |
da8be7fabf5e4376921a77b2c3cf7b7f
|
Constraint plane S: (i_1+i_2)/2 + h = x.
|
cond
|
mat0408260-fig4.jpg
|
0.573238 |
f70fc0929c794bcb8c3411ba02b908b6
|
Comparison of simulation times for the three algorithms discussed in the text. Circles are used for the lattice-model algorithm, squares for the double-list algorithm, and triangles for the phantom-nuclei algorithm. For each system size, the number of Monte Carlo realizations ranges from 5–20, and the lines connect the average simulation times. The double-list algorithm is two to three orders of magnitude faster than the lattice algorithm, while the phantom-nuclei algorithm ranges from three to five orders of magnitude faster, depending on the number of time points at which one records data. The filled triangles show the fastest case, with only one time point requested, while the open triangles show the slowest case, where data are recorded at each intermediate time step.
|
cond
|
mat0408260-fig5.jpg
|
0.478773 |
d03fa6e2d9d54c41b747b93fa5aef618
|
Schematic description of the double-list algorithm. (a) Basic set-up for lists {i} and {h'}. Note that {h'} records cumulative lengths. (b) Nucleation. (c) Coalescence due to growth.
|
cond
|
mat0408260-fig6.jpg
|
0.433299 |
6c9e2169f8164962b19ca33f1f0d58d2
|
Schematic description of the phantom-nuclei algorithm. The figure shows the distribution of potential initiation sites in the spacetime plane. The open circles denote sites that do initiate, while the “phantom" filled circles, lying in the “shadow" of the open circles, do not initiate.
|
cond
|
mat0408260-fig7.jpg
|
0.449792 |
ff576b1c362c4554b2b5ca46b3df69ba
|
(Color online). Theory and simulation results for I(t) ∼ t. Size distributions are calculated at these timepoints: t = 50, 75, and 100. (a) Hole-size distribution ρ_h(x,t). (b) Island distribution ρ_i(x,t). The inset plots f(t) vs. t, with the dot at t=75 (f=0.5). (c) Island-to-island distribution ρ_i2i(x,t). The analytical curves have been obtained by Eq. <ref>. There is a crossover of the decay constant slightly after t=75 (f=0.5) (see text). The inset shows ρ_h(x,t) and ρ_i2i(x,t) for t=50. All figures have the same vertical range of 10^-7 - 10^-3 (log-scale).
|
cond
|
mat0408260-fig8.jpg
|
0.475907 |
7c3646fef870431d8cb83f83c67b81f7
|
(Color online). Decays constants of ρ_h(x,t), ρ_i(x,t), and ρ_i2i(x,t). The symbols are simulations, and the solid lines are theory.
|
cond
|
mat0408260-fig9.jpg
|
0.453836 |
da39f76105a74c69987d302552123bb3
|
The schematic phase diagram of the dilute Bose gas with single Feshbach resonance near the transition line between the ASF and MSF phases below T_c(δ), where T_c(δ) represents the transition temperature between the normal Bose gas and the superfluid phase. The dotted line denotes the crossover from low temperature to high temperature regimes. The solid line is the phase boundary between the ASF and MSF phases. The point C is a tricritical point. On the phase boundary, the portion below the point C is of first order while the one above the point C is of second order.
|
cond
|
mat0408267-bfphase.jpg
|
0.446165 |
ce88bee7b4f34d8ca5c2d44f6ffa1abf
|
Typical solutions of Eqs. (<ref>)—(<ref>) with the initial conditions (a) r=0.002, K=0.0003, and U=0.005, (b) r=0.002, K=0.003, and U=0.005, and (c) r=0.002, K=0.03, and U=0.005. The x-axis is r_l and the y-axis is K_l/K or U_l/U.
|
cond
|
mat0408267-rgfig1a.jpg
|
0.489948 |
a3432d7bc271459baa5bc4e667764159
|
Peierls instability in a lattice fermi-bose mixture: (a) repulsive fermion-boson interactions (b) attractive fermion-boson interactions. The shaded part depicts the bosonic mean field density whereas filled circles denote fermionic atoms.
|
cond
|
mat0408269-hol1.jpg
|
0.428572 |
66b3bcb1032b4220b4c0fc786f55e262
|
Fermionic energy spectrum for various values of the bosonic modulation wavenumber with ω_0=0 (a) and as a function of ω_0 for a fixed modulation with k/M=1/2 (b).
|
cond
|
mat0408269-hol2.jpg
|
0.566358 |
3a220a15ad78498a9acdf56cfb9a6e0d
|
Ground-state fermionic energy as a function of the modulation wavenumber for various fermion filling factors: N_c/M=1/2 (solid), N_c/M=1/4 (dashed), N_c/M=1/8 (dash-dotted). The external trap frequency is set to κ=0.1t_c and the bosonic amplitude modulation is set equal to t_c. Arrows indicate the bosonic modulation wavenumber minimizing E_c.
|
cond
|
mat0408269-hol3.jpg
|
0.553633 |
09ba618bf533464b8d330b0c03ac92b2
|
Total energy of the system as a function of the gap Δ for various values of t_c (a) and optimal gap values as a function of (b). The dashed line depicts the predicted strong coupling behavior of Eq. (<ref>).
|
cond
|
mat0408269-hol4.jpg
|
0.473801 |
0e7649ed373f46c3917c2b8106c3fb76
|
Scattering in details, see previous figure
|
cond
|
mat0408271-Fig10.jpg
|
0.484249 |
48251a26341d452290aee2756d615f0b
|
Solitons in difference scheme <ref>, c=-i, λ=0.5
|
cond
|
mat0408271-Fig11.jpg
|
0.431697 |
8c0aa1f6fae440d2ba85214d489004fe
|
Velocities of solitons, which occurs in the limits of integration range and in its center, are equal
|
cond
|
mat0408271-Fig12.jpg
|
0.447734 |
56a5c25edb474130b821ab54b35cbee5
|
“Soliton-type“ tunneling
|
cond
|
mat0408271-Fig13.jpg
|
0.514658 |
f4db39fe0f92412ebabd51a333946e5e
|
“Tunneling“ rate for λ=1. Origin of the plateau corresponds with escape of soliton 3 from under the barrier; ending of the plateau corresponds with entering of soliton 3 under the barrier
|
cond
|
mat0408271-Fig14.jpg
|
0.50991 |
3effd98b3707492480b9159898d78c14
|
Solution for the linear diffusion equation. c=-1, i ∈[1,500], j ∈[1,30]
|
cond
|
mat0408271-Fig2.jpg
|
0.488575 |
88cdf35fc12a4971b60955fbf7afd85e
|
Solution for the linear Schrödinger equation, c=-i
|
cond
|
mat0408271-Fig3.jpg
|
0.495924 |
8dc12acadfa84b5d8181f185a4e4952e
|
Solution for the difference scheme with parameters c=-1, λ=2.1
|
cond
|
mat0408271-Fig4.jpg
|
0.472746 |
190e9916f1df474aabe3de5759fb66f4
|
Solution for the difference scheme with imaginary parameter c=i, λ=2.1
|
cond
|
mat0408271-Fig5.jpg
|
0.469472 |
7ab8bb0cff7f43e4a50a358e1e719aa3
|
Divergent solution for the difference scheme with c=-1, λ=1.9
|
cond
|
mat0408271-Fig6.jpg
|
0.486467 |
4f0c7cd12159445d9461bfd3f6f2f9c7
|
Soliton-like solution for c=-i, λ=1.9
|
cond
|
mat0408271-Fig7.jpg
|
0.464288 |
ddecf49faa0d454ebe308f961d0291be
|
Solitons for c=-i, λ=1
|
cond
|
mat0408271-Fig8.jpg
|
0.475929 |
3f9220a9872c4f34bf1b1fe5df749201
|
Scattering of solitons, c=-i, λ=1, m=350
|
cond
|
mat0408271-Fig9.jpg
|
0.482323 |
d582e4d3b73248aea2752947521be2d2
|
Conductivity vs. temperature, calculated for wide edge state with parameters chosen to reproduce experimental results of Ref.<cit.>. Inset: results for narrow edge states with interaction strengths κ a=1 (∘) and κ a=5 (♢).
|
cond
|
mat0408272-fig1.jpg
|
0.583642 |
b8001f11284142d0bc4da8c5df92eda5
|
The scaling function C(δ B/B_0) for conductance fluctuations: for narrow edges with κ a=0.6 (∘) and κ a=1 (); and for wide edges with the parameters of Fig. <ref> ().
|
cond
|
mat0408272-fig2.jpg
|
0.547752 |
cdc647f47f994f63b15b7d899847a117
|
S/F system. The ferromagnet shows a type I BS. The DoS at the Fermi level for majority band (spin up) is larger than the DoS of the minority band. The two electrons connected by a dashed line represent a Cooper pair which contributes to the inverse proximity effect.
|
cond
|
mat0408273-fig1.jpg
|
0.486152 |
9018818d381048a4a357c463b90a8092
|
S/F system. The ferromagnet shows a type II BS. The DoS at the Fermi level for minority band is larger than the DoS of the majority band.
|
cond
|
mat0408273-fig2.jpg
|
0.478363 |
caad7b4793694d718a2fe3273187ecd2
|
S/F structure consisting of a ferromagnet with a large exchange splitting h. The band with of the minority spin-band is approximately equal to the band width of the superconductor.
|
cond
|
mat0408273-fig3.jpg
|
0.461626 |
420278ac0e6f41c498a2c0a4f8d93482
|
S/F systems. The magnetization in F is due to the magnetic moment of certain regions. The screening or antiscreening in S is only possible in regions of size ξ_s (circles)
|
cond
|
mat0408273-fig4.jpg
|
0.453931 |
9263cf10339a460bb99a3a6fa7a71d23
|
(a) PL image of sample with current I=+200 μA at B=3 T. PL is seen as light gray, black areas are the metallic contacts. Central circular area is currentless. (b) Difference of (a) and the PL image of a currentless sample. Contact edges of the Corbino sample under current (otherwise invisible in a difference image) are marked with dotted white lines.
|
cond
|
mat0408274-fig1.jpg
|
0.437674 |
201e7cd5f05d4fa9986ee9f19a2a4881
|
(a) Construction of streak images from series of sample PL images at varying B. Black vertical balks indicate sections in the currentless (upper balk) and in the current biased (bottom balk) parts of the sample. Vertical dimension of both sections is 400 μm. (b) Streak image of the sample section without current. Relative PL intensity is color coded according to palette on the right. (c) Streak image of the sample section with current bias I=100 μA. The fan of dotted eye-guides indicates the dark structures (rings) travelling from the positive to the negative contact in increasing magnetic field, in contrast to standing oscillations in Fig.(b). (d) Electrically measured oscillations of the sample voltage. Coincidence of the voltage maxima and the dark structure extinctions at the negative contact is indicated by the vertical dotted lines.
|
cond
|
mat0408274-fig2.jpg
|
0.484829 |
64a56a37d7d44d3ea995203bd153339e
|
(a) 2DEG density profiles evaluated from the 1/B periodicity of the streak image in Fig. <ref>c. Triangles and squares correspond to I=+100 μA and I=-100 μA, respectively, with higher n_2D always at the negative electrode. Solid line indicates the zero-current 2DEG density acquired from Fig. <ref>b. (b) 2DEG density evaluated from oscillations of sample voltage. Full and open circles correspond to sample before and after illumination, respectively.
|
cond
|
mat0408274-fig3.jpg
|
0.444772 |
333e4713a7ea42bb8cf538bea2e09e04
|
(a) Schematic sketch of Fermi (dashed line) and Landau levels (solid lines) in one electron approximation. Arrows indicate direction of crossing point motion in increasing B. (b) Sketch of Fermi and Landau levels with electron screening. Compressible stripes are indicated by dotted balks.
|
cond
|
mat0408274-fig4.jpg
|
0.434549 |
2da38fdf784144b2b69628c0d7eff60e
|
(a) Differential conductance g (log color scale) as a function of source-drain bias V_SD and gate voltage V_G at B_⊥=0, at base electron temperature T_el=45 mK. Numbers 0 through 4 are number of electrons in the dot. White vertical lines identify the locations for data shown in (c) and (d). (b) Same as (a), at B_⊥=1 T. (c) Differential conductance through the N=2 diamond showing step with overshoot at V_SD = J(B_)/e at B_=0 and 1 T. (d) Differential conductance through the N=3 diamond showing Kondo peak at V_SD = 0 for B=0, split by B_⊥=1 T.
|
cond
|
mat0408276-fig1ST9.jpg
|
0.475837 |
9069a244c90a4afead49d42d5ddf7fbf
|
(a) Differential conductance g as a function of V_SD in the N=1 diamond (V_G=0.1 V) for in-plane fields B_X=0,0.4,0.6,0.8,1 T, (top to bottom, curves offset). Dashed grey lines are guides to the eye showing the cotunneling gap. (b) g(V_SD) shows a zero-bias peak in the N=3 valley (V_G=0.42 V) that splits in an in-plane field B_Y=0,0.25,0.45,0.7,0.95 T (top to bottom, curves offset). (c,d) splitting energies (see text) versus magnetic field as in (a,b) with linear fits. Insets: angular dependence of the g-factor in the plane of the 2DEG indicating isotropic behavior. Dashed circles show direction-averaged g-factors. Directions X and Y in the plane are arbitrary.
|
cond
|
mat0408276-fig2ST9.jpg
|
0.470289 |
55ae6f32cbc8457789c158b0df85c04a
|
(a) Differential conductance g (log color scale) as a function of V_SD and V_G for B=0 T in the vicinity of the N=1→2 transition. (b) g(V_SD, B_) at V_G=0.164 V (vertical white line in (a)) <cit.> shows the perpendicular field dependence of the singlet-triplet gap. (c) Cuts showing g(V_SD) at the positions of the vertical lines in (b), marked A, B, C, D. (d) Horizontal cut (E in (b)) showing g(B_) at zero bias. Note the asymmetric peak in g at the singlet-triplet transition.
|
cond
|
mat0408276-fig3ST9.jpg
|
0.488086 |
418d6a437bfc4274be93bf182663857c
|
(a) First and second one-electron excited state energies Δ_1 and Δ_2, measured from sequential tunneling along with fits to a 2D anisotropic harmonic oscillator model with ħω_a=1.2 meV and ħω_b= 3.3 meV (see text). (b) Singlet-triplet splitting J as a function of magnetic field B_. (c) Dependence of J on gate voltage V_G for various B_ as indicated. (d) Average slopes Δ J(V_G)/ Δ V_G from (c) as a function of magnetic field B_, showing strong reduction of gate voltage dependence of J at large B_.
|
cond
|
mat0408276-fig4ST9.jpg
|
0.488559 |
aec220833e38480d863ec03560fa70d7
|
Differential conductance g as a function of V_SD for temperatures T_el=45, 140, 170, 240, 280, 330, 380, 420, 570 mK (top to bottom) showing overshoot at V_SD∼ J/e. Inset: peak conductance as a function of temperature with best-fit log(T) dependence (line).
|
cond
|
mat0408276-fig5ST9.jpg
|
0.451232 |
2791c665218245998282a781a17b36ab
|
Magnetic contribution ρ_mag(T) to the electrical resistivity of CeRu_2Ge_2 at different pressures. Two different antiferromagnetic phases occur below T_N and T_L, and a ferromagnetic ground state is present below T_C and low pressure. No traces of magnetic order are observed for p>7 GPa above 1.2 K. Inset: Pressure dependence of the A-coefficient obtained from a fit of ρ(T)=ρ_0 + AT^2 for T< 0.5 K to the data of Ref. <cit.>.
|
cond
|
mat0408280-wilhelm_fig1.jpg
|
0.457319 |
0ee1815a195b4e0289bd9ca05382562e
|
Plots showing a_0(x,t_0) vs a_1(x,t_0) (left panel) and a_0(x_0,t) vs a_1(x_0,t) (right panel) for Periodic [(a),(e)], Chaotic [(b),(f)], (C→A) [(c),(g)] and aligned [(d),(h)] regimes.
|
cond
|
mat0408282-SPphaseportraits.jpg
|
0.36739 |
c16d74b7852c407bbc1aee5f3d29d011
|
(Color online) Space-time evolution of the shear stress showing co-existence of different dynamical regimes at γ̇=3.9,λ=1.12.
|
cond
|
mat0408282-coexistSTOandT.jpg
|
0.479892 |
f4c4bf79817d4c5ca0d8af2fa9a84793
|
Decay of the distribution of laminar domains at the onset of the spatiotemporal intermittent regime (represented by circles) on Log-Log scale. The solid line is a power law fit to the data with an exponent ψ_STI=1.86 ± 0.05.
|
cond
|
mat0408282-distlaminardomains.jpg
|
0.434603 |
b81f98aa7362439383770744fdb4cb49
|
Decay of the distribution of laminar domains away from the above-mentioned onset and well within the STI regime (represented by circles) on a semiLog scale.
|
cond
|
mat0408282-distlaminardomainsexpgdot4.jpg
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.