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
|
---|---|---|---|---|
0.506329 |
ef28484d1b124e4cb5a91945a636686a
|
Simulation for α = 0.7.
|
2012.09317
|
path_alpha07.jpg
|
0.456111 |
1fb27033b5da40b8813830cc1b6bab34
|
Simulation for α = 0.8.
|
2012.09317
|
path_alpha08.jpg
|
0.558488 |
1c21a41e86944ab1b5ca97de2376eb45
|
Simulation for α = 0.9.
|
2012.09317
|
path_alpha09.jpg
|
0.443299 |
a734c812238944658cfbf31c39e28f89
|
Simulation for α = 1.
|
2012.09317
|
path_alpha10.jpg
|
0.436989 |
3d04fe5e80e44497a6b2a1703727b82d
|
Summary of two types of process which occur in the brain and seem to map onto Kahneman's System I and System II.
|
2012.09318
|
two_types_of_Process.jpg
|
0.487699 |
4116ba1c186341849dda5d6b37dcd615
|
A Comparison of the absolute bias of the R-estimator with its standard error
|
2012.09445
|
Figure0.jpg
|
0.467319 |
7eb7ecf4f608433086a919dfed32e889
|
Estimate of the Tail Parameter in 1,000 Years of Simulated Monthly Data
|
2012.09445
|
Figure2.jpg
|
0.412265 |
40d42e46873b44e38ae31ec35e4463d3
|
The Lorenz Curve for a Single Simulation
|
2012.09445
|
Figure3.jpg
|
0.434534 |
9bad90e7230640f1b9d97fe52571cec6
|
300 Years of Simulated Weekly Data in an Economy with One Million People
|
2012.09445
|
Figure5.jpg
|
0.529833 |
581961dcc5844a439d58df7df7880baa
|
The Evolution of Average Beliefs When Agents are Least-Squares Learners with Infinite Lives
|
2012.09445
|
Picture1.jpg
|
0.474894 |
1ea87597fece47dc8a3b8128ec240a44
|
The Evolution of Average Beliefs When Agents are Constant-Gain Learners with Finite Lives
|
2012.09445
|
Picture3.jpg
|
0.455372 |
8a1dc28f81ff49a6b01f1dbe92145802
|
ALMA Band-6 images before tapering of 12 targets (image size: 5”×5”). A black circle shows the beam size of each image. Black contours correspond to 4σ, 8σ, 12σ, and 16σ. A plus mark represents the centroid determined in the Ks-band image. Some of the ALMA detected sources, such as 406444 and 434585, show a larger spatial offset between the dust continuum emission and the Ks-band centroid than the other detected sources. This is probably due to the lower signal-to-noise ratios of their dust continuum emission. The stacked image of the five individually ALMA non-detected sources with log(M_*/M_⊙)= 10.0–10.4 is also shown. The central position is shown with a plus mark. The stacked emission is detected at 5σ.
|
2012.09447
|
figure1.jpg
|
0.487482 |
9329a830cb564b71900bcdb7c73532dd
|
Gas mass fraction versus gas-phase metallicity for our galaxy sample at z∼3.3. The horizontal bar in the bottom right corner shows an additional ± 1σ error on the gas mass fraction coming from the systematic uncertainty on M_gas. The dashed lines show how the two quantities depend on each other when fixing the dust-to-stellar mass ratio at 1×10^-3 and 3×10^-3. We find no statistically significant correlation between the gas mass fraction and gas-phase metallicity among our sample.
|
2012.09447
|
figure5.jpg
|
0.542845 |
a6c016e978e14b5a8c3592b377cd6668
|
Relation between gas mass fraction and gas-phase metallicity for star-forming galaxies from z=0 to z∼3.3. The horizontal bar in the left bottom corner represents an additional ± 1σ error on the gas mass fraction coming from the systematic uncertainty on M_gas for our sample. Only the molecular gas components are considered for the galaxies in <cit.> and <cit.>. The star-forming galaxies at z≳2 show an offset toward the lower gas-phase metallicity from the distribution of the local galaxies. The black lines show the model tracks from the gas regulator model of <cit.> assuming different mass-loading factors between λ= 0.5 and 2.5. The distribution of the star-forming galaxies at z∼3.3 on this diagram can be broadly explained with the model tracks with λ∼ 2–2.5, suggesting the redshift evolution of the mass-loading factor for star-forming galaxies.
|
2012.09447
|
figure6.jpg
|
0.488328 |
8ce8df4de8b24939a2c2687b91746299
|
Best-fit SEDs (sold line) obtained with magphys. Data points are from the COSMOS2015 catalog <cit.> and the ALMA Band-6 observation in this study. Arrows show 3σ upper limits. When fitting the ALMA non-detected sources, the 1.3 mm flux and its uncertainty are set to 1.5σ± 1σ <cit.>.
|
2012.09447
|
figure7.jpg
|
0.409336 |
e5f0782c95ff442486ce529844cacb46
|
Frequency histogram of weighted average of relative error of estimated ATE for eight different models: ordinary least square(OLS), LASSO, RIDGE, random forest (RF), GRU, CNN, and multilayer perception (MLP) based on the semi-synthetic data with N=160000, M=500, and α=0.05 and β=0.05 in (<ref>) of the main paper.
|
2012.09448
|
ATE_JD_160000_500_8_model.jpg
|
0.434653 |
efcf44aa6e5c4297b680f196adbff6bf
|
Weighted average of relative err. of estimated ATE using IoC and IwC for different models; N=160000, M=500, and α=0.05 and β=0.05 in (<ref>).
|
2012.09448
|
ATE_JD_160000_500_new.jpg
|
0.446368 |
f31af8a39de34973922c0d12c7cdffbd
|
Frequency histogram of weighted average of relative error of estimated ATTE for eight different models: ordinary least square(OLS), LASSO, RIDGE, random forest (RF), GRU, CNN, and multilayer perception (MLP) based on the semi-synthetic data with N=160000, M=500, and α=0.05 and β=0.05 in (<ref>) of the main paper.
|
2012.09448
|
ATTE_JD_160000_500_8_model.jpg
|
0.464055 |
a80322035b304afa88407869b05ec908
|
Weighted average of relative err. of estimated ATTE using IoC and IwC for different models; N=160000, M=500, and α=0.05 and β=0.05 in (<ref>).
|
2012.09448
|
ATTE_JD_160000_500_new.jpg
|
0.465121 |
2a44c23537b44e12a855b8c87becfae0
|
The mean relative error computed by (<ref>) vs. number of observations.
|
2012.09448
|
simu_cond_exp_consistency.jpg
|
0.402974 |
5b7bdf1b96d548718ed8f9ea60418717
|
The average standard deviation computed by (<ref>) vs. number of observations.
|
2012.09448
|
simu_cond_exp_std_consistency.jpg
|
0.428618 |
8b5d93c8b0d042daaa039b9c6f6d80b5
|
The mean relative error computed by (<ref>) vs. number of observations.
|
2012.09448
|
simu_exp_convergence.jpg
|
0.377465 |
0c97092d9dd34d8488c6aceaf2697dba
|
The average standard deviation computed by (<ref>) vs. number of observations.
|
2012.09448
|
simu_exp_std_convergence.jpg
|
0.422849 |
26ad506c557948ea9181629cd8b048b9
|
A photo of the MAFDS and its experimental test setup.
|
2012.09449
|
DSC_2177_opt_ausschnitt_cut_grau.jpg
|
0.512908 |
30786296a11c4746922a636b779d4a91
|
Plot compares density estimator of piezo-beam, cf. Subsection <ref>. A kernel density estimator based on all available experimental data (gray solid line), a kernel density estimator based on the computer model (black dashed line) and kernel density estimator based on an improved surrogate model (black solid line). The values of the outcome of the 10 experiments are indicated on the x-axis.
|
2012.09449
|
KeKo2019_PIEZO_plot.jpg
|
0.472585 |
6f923541b9f2485b8996dff46035aff3
|
Box plot without outliers of 10^6 estimates for the 0.95-quantiles of the n=100 absolute model errors for both computer models of the MAFDS.
|
2012.09449
|
KeOh2001_MAFDS_plot.jpg
|
0.498739 |
9df68f32baac46969c3bcc27fb7626e5
|
Three estimates of the cumulative density function of the experimental outcome of the MAFDS. An estimator based on experimental data (black solid line), an estimator based on computer experiments with computer model 1 (black dashed line) and computer model 2 (gray dashed line).
|
2012.09449
|
RoOb2011_MADFS_plot.jpg
|
0.476981 |
e0d5531ccdc745cab3cf608526334c82
|
Box plot without outliers for the 0.95-quantiles of the n=100 absolute model errors for both computer models of the MAFDS. The box plot contains B = 500 bootstrap 0.95-quantiles.
|
2012.09449
|
WoStLe2017_MAFDS_plot.jpg
|
0.484634 |
3de45e8e2794425c8848830ef350c1f4
|
A CAD model of the lateral vibration attenuation system with piezo-elastic supports and a sectional view of one of the piezo-elastic supports, cf. <cit.>.
|
2012.09449
|
piezo_beam.jpg
|
0.421755 |
6674ae92ed4745bbb0ba484a8d6db170
|
LS-F.
|
2012.09451
|
1872_F-eps-converted-to.jpg
|
0.416188 |
59f46f7791a84eeeb49c73e6a51f6dcc
|
LS-G.
|
2012.09451
|
1872_G-eps-converted-to.jpg
|
0.469548 |
d1fec0fc822b4cc9869ff97bca2f431f
|
The results with different values of k.
|
2012.09451
|
big_k_trim-eps-converted-to.jpg
|
0.444253 |
752ecb6225a04c2b8256ea27a2c5cf42
|
The number of blocks before and after applying LS-G and LS-F on METIS.
|
2012.09451
|
block_trim-eps-converted-to.jpg
|
0.459049 |
dda1c23b70a84dda95b129d4185ca50a
|
Predictions for the mass of the core and the amount of metallic elements in the envelope obtained using the 3rd order theory of figure. The published results are compared with the ones obtained by varying the d_0 parameter are shown in the figure and for a core constituted of pure water and of silicate and water.
|
2012.09454
|
fig6-c-arxiv.jpg
|
0.423496 |
8b847ac162e445a0bf4be4da6e45314a
|
Optical plasma diagnostics show strong self emission from dense narrow regions of the plasma. We see (a) raw interferogram measurement of the plasma, 1 ns after the interaction starts, measured with the Nomarsky interferometer (b) the retrieved electronic density of the plasma and (c) the raw streak camera measurement of the self emission of the same laser shot (including measurement lineout in white). x=0,y=0 are the coordinates of the jet shock point; the laser travels left to right and focalizes in the same point.
|
2012.09455
|
Image_worm.jpg
|
0.444622 |
8b0d461201064cf5bfa9eadf0ec8cb8a
|
RCF imprints converted to dose and alpha particle number density for three consecutive shots varying the gas jet density profile. The number density conversion presumes the appearance of alpha particles, a cut-off energy equal to the last signal layer and a step-wise flat spectral shape between layers. The typical FWHM divergence angle of the beam imprint is 9, all images have the same spatial scale. Note that number densities are normalized to one pixel, with respect to the resolution of RCF scans of 600 dpi. For shots #62 and #63, the presumed pre-aligned laser axis corresponds to the centre of the illustrated frames. For shot #64 this does not apply.
|
2012.09455
|
RCF_62_63_64_data.jpg
|
0.498145 |
b9060f6aeeaf415896fc87f94f3d2afc
|
Projectile spectrum for three consecutive shots varying the gas jet density profile, presuming an alpha particle beam impacting on the stack of RCF. RCF color channels R, G and B are independently analyzed.
|
2012.09455
|
RCF_62_63_64_spectrum.jpg
|
0.407792 |
0687f70542944dca8c54315cacff4f89
|
Projectile spectrum for three consecutive shots varying the gas jet density profile and transverse position, presuming an alpha particle beam impacting on the stack of RCF. RCF color channels R, G and B are independently analyzed.
|
2012.09455
|
RCF_79_85_87_spectrum.jpg
|
0.445226 |
a0c847948322455fa1ff670d95909772
|
Electronic temperature chart in MeV of the n_at,max = 4× 10^19cm^-3 (0.036 n_c) simulation.
|
2012.09455
|
Te_TUCAN.jpg
|
0.458839 |
f8093faaaebf4b37b436f9a1f8cc5337
|
Diameter evolution of an individual etch pit on a TasTrak CR-39 with strioscopic snapshots at 2, 3, 9 and 15 of etching with 6.25 N etchant at 70. The diameter evolves considerably and is not detected for 2 of etching as it may be underneath the resolution of the microscope system.
|
2012.09455
|
centerion_pit_strio_sketch.jpg
|
0.396265 |
687afba797c34bd9b54809575e447ff1
|
Summary of results.
|
2012.09455
|
comp.jpg
|
0.429131 |
24ed3b27bb40441fa37c7c8f1a1be7e0
|
Comparison between the detector RAW signal comprising photopeak and all accelerated particle species (blue full line) and a signal with photopeak only (orange dashed line).
|
2012.09455
|
comp_photopeak_norm.jpg
|
0.382355 |
5463e70af6cf41d0b9206be144dbc96e
|
Photodiode detector impulse response (left) and signal after deconvolution (right).
|
2012.09455
|
deconv_subplot.jpg
|
0.426681 |
228c3dbf390649698a2ffb08618323d9
|
E_y field chart resulting from the simulation with a density profile with n_at,max = 4× 10^20cm^-3. The blue dashed line marks the density peak.
|
2012.09455
|
ey_TUCAN.jpg
|
0.389623 |
fbf45ad50ee140e4962492ad934ed9b2
|
(a) Interferometric images overlaid by self-emission line-outs for shots with variation of the transverse position of the laser-gas interaction. The longest plasma channel is recorded for shot #81 with positive transverse displacement – the length of the channel decreases for the series of shots towards negative transverse displacements. The last shot of the series #83 does not agree with this assessment and shows a longer channel again. Note that (b) un-driven gas density profiles acquired prior to corresponding shots show visual changes from shot-to-shot, indicating changes to the nozzle. This agrees with a qualitative difference of the nozzle surface between shot #77 and #83, presented with fig. <ref>.
|
2012.09455
|
gas_77_80_81_82_83.jpg
|
0.44898 |
96bdf814c360425e8b0569270869b82f
|
Shot-to-shot variation of the gas jet density profile using a mix of Helium and Nitrogen in a ratio 1 to 9, (top row) longitudinal un-driven gas jet density profile, and (bottom row) interferometric image superposed to aligned laser beam and plasma self-emission in arbitrary units as red line, 1ns after the interaction starts. The laser is focused to x=0,y=0, the coordinates of the jet shock point, but transverse displaced by 0;±50.
|
2012.09455
|
gas_79_85_87_overview.jpg
|
0.507553 |
138f4487b1674706b8dd157debac2d39
|
Diameter evolution for similar groups of etch pits on a TasTrak CR-39 with snapshots at 2, and 9 of etching with 6.25 N etchant at 70. Highlighted are groups (a,b,d,e) with circles concentric to the etch pit position for the short etching time and (c) with circles concentric to the pit position for long etching time. The dashed line delimits a scratched area, we do not see the edge of the plastic sheet.
|
2012.09455
|
monstermark_pit_brightfield_sketch.jpg
|
0.41038 |
b0ac722be9ce44bbac75db5749106150
|
Overview on RCF results for shots #79, #85 and #87 converted from RAW to deposited dose maps.
|
2012.09455
|
qulitative_RCF_stack_Al4UEBT32EBT3.jpg
|
0.414494 |
53ee9e3a33954c2ba9449350b06cd95b
|
Nitrogen ions density chart (left) and zoom on one of the ion acoustic waves travelling in the down-ramp of the density profile (right) for the first simulation run with a_0 = 6.8.
|
2012.09455
|
qx_TUCAN_and_zoom.jpg
|
0.496641 |
2e9c4e4dbbf449f6adc6c707b2a5c8a1
|
Z-scan of etched pits on a TasTrak CR-39 after 15 of etching with 6.25 N NaOH etchant at 70. The bulk etching rate was quantified to 1.32. Experimental data points are underlaid with a relative point number density color map in grayscale. The simulated spectrally resolving species specific L(D) lobes are for alpha particles on the left hand side and for fully ionized Nitrogen ions on the right hand side. Specific impact energies are highlighted with red marker arrows (1–4). Experimental data points that overlap with the lobes in the range of their uncertainty are denoted with the closest corresponding energy on the lobe. Simulated with the CR-39 plug-in of PySTarT.
|
2012.09455
|
track_diameter_energy_15.jpg
|
0.393433 |
f384dd6127914218b1a195583a6d9dad
|
An analytical investigation of the critical transmissibilities of both strains for increasing clustering coefficients for the ER networks defined in Fig <ref>. The curves are generated from Eqs <ref> and <ref> at the value of T in which the outbreak fraction becomes large than ϵ. The critical points reduce as C increases, indicating that contact clustering helps to spread an emergent epidemic. The interval [T_1^*,T_2^*] is the transmissibility window in which strain-1 exists as an epidemic on the network without strain-2. We observe that increased clustering reduces the interval and thus increases the extent of coinfection in this model, at the expense of the mono-infection equilibrium. We note, however, that clustering can never reduce the coinfection critical point below the single strain threshold due to the strict conditions of perfect coinfection in the premise of the model.
|
2012.09457
|
Delta_transmissibilities.jpg
|
0.406831 |
cf355fd736d649b9b393e7f911db2b9a
|
(Top) The outbreak fractions of both strains for increasing triangle probabilities within the CHCN degree distribution defined in Eq <ref>. Clustering reduces the epidemic threshold, in agreement with ER experiments; however, coinfection decreases with increasing θ. (Bottom) The expected epidemic sizes of both strains for a p^CHCN(k) network with θ=0.0 and a degree-δ CHCN distribution network defined by Eq <ref>. Clustering increases the epidemic threshold in this model. Markers are the average of 100 repetitions of bond percolation on CHCN networks with N=1e^5 nodes and T_2=0.6.
|
2012.09457
|
PerfCoinfFig5PlotB.jpg
|
0.428417 |
9e21e7b4c97a421984bbdb01424cc315
|
The tree-like edge topologies found in the GCC following the first strain and their probabilities.
|
2012.09457
|
TREES_coinf.jpg
|
0.449054 |
32d42d02a2424f7eb789ca98b5ea1df0
|
The 3-cycle edge-topologies found in the GCC following the first strain and their probabilities. Node colours and patters are defined according to Fig <ref>. Triangles D, E and F consist of inhomogeneous neighbour-states and hence, due to the symmetry of the shape, their reflection about a vertical axis bisecting the focal node can also occur with equal probability. The curved arrows in triangles D and F indicate the additional pathway through the cycle that the infected neighbour could infect the focal node.
|
2012.09457
|
TRIANGLES_coinf.jpg
|
0.541495 |
ae746598d94a4951a87ab0f3d011efa2
|
The outbreak fractions for the 2-strain coinfection model on ER random graphs with T_2=0.6, N = 25000, for varying clustering coefficients and fixed mean degree ⟨ s⟩+2⟨ t ⟩=2. It is clear that clustering reduces the epidemic thresholds of both diseases and reduces the outbreak size of the first strain; however, it increases the coinfected fraction of the network. Points are the average of 500 repeats of Monte Carlo simulation while solid lines are the theoretical predictions from Eqs <ref> and <ref>.
|
2012.09457
|
coinfection.jpg
|
0.443824 |
eb01c3e963fd431a9eae58660704fde1
|
The expected number of bars of length ≥, N^_B, as a function of .
|
2012.09459
|
BMCalculated.jpg
|
0.445585 |
e5e90015606e4f589e11585d2272602b
|
N^_B as approximated with simulations of random Rademacher walks.
|
2012.09459
|
NvepsBMnothing.jpg
|
0.405588 |
c40d5ac8926d49b8b25d83705a18a0cc
|
A function f: [0,1] → and its associated tree T_f in dashed lines.
|
2012.09459
|
TreeFunction.jpg
|
0.433145 |
65a32d573bf44ee38c0d1d3885863f75
|
A depiction of the first steps of the algorithm which assigns a barcode (f) to a tree T_f.
|
2012.09459
|
algorithm.jpg
|
0.433826 |
1a2dcfb0306f4ad3a89510733afe1eda
|
Comparison between theoretical predictions and finite element simulation for twist (a) and stretch (b) as a function of the spontaneous deformation parameter λ_s. For these calcuations, we used a fibre with aspect ratio R_0/L=1/15 represented by 5400 hex-8 elements.
|
2012.09461
|
TwistStretchComparison.jpg
|
0.451198 |
4e80121ea0574c56bf57822187014d96
|
Finite element simulation of a fibre with Young modulus E=1, Poisson ratio ν=0.45, 4800 elements, aspect ratio of R_0/L=0.07 and a R independent director-field with α=π/4 and β=π/2. The fibre is fixed at one end and subject to a tension T=0.2 πμ R_0^2 while preventing rotation at the other end. As the value of λ_s is decreased, the fibre wants to shrink and twist but cannot do the latter as the ends are not allowed to rotate. The fibre thus coils to release some of the twist energy, resulting in a greater stroke amplitude then that of classical linear actuation.
|
2012.09461
|
coil.jpg
|
0.40009 |
2a9c956b6e814e7bb1dfecf4b3f8bfa1
|
Comparison between Finite element simulations and theoretical data for the case α=π/4 and β=π/4. a) An example of a deformed cross section directly from FE analysis. In red are theoretical lines for the deformed radii while dots are node positions. b) Comparison between predicted and theoretical coning. c) Comparison between theory and FE on the predicted rotation angle difference as a function of radius. A fibre with aspect ratio R_0/L=1/5, and 54000 hex-8 elements was used for this simulation.
|
2012.09461
|
coning2.jpg
|
0.407857 |
4d702e0350634fe4b7804395911f467c
|
a) 2D plot of twist as a function of both angles α and β. The twist is maximised when β=π/2 and α∼π/4. b) – c) Variation of the twist and stretch as a function of α for different values of λ_s.
|
2012.09461
|
costant2.jpg
|
0.401871 |
543a1573c7c14894a3347b74cded426a
|
Relationship between output twist a) and stretch b) as a function of the twist imposed at genesis. Shaded in grey, the region for known experimental results <cit.>.
|
2012.09461
|
genesis.jpg
|
0.444126 |
8bb867071cb64e048c5453869f4940bb
|
a) Schematic of a fibre being drawn and twisted from a drop of LC monomer while being cured with UV light. b) An example of the resulting director field in a fibre of radius r=2/τ_0. The two integral curves highlight the change in azimuthal component as a function of the radius.
|
2012.09461
|
genesis2.jpg
|
0.419085 |
1e886a40c1fe412da0ec9f645961dc06
|
(a) The optical dipole potential energy for right-handed circular polarization. Solid line represents the case where ℓ=+1, dashed curve represents the case where ℓ=-1; (b) The optical dipole potential energy for left-handed circular polarization. Solid line represents the case where ℓ=-1, dashed curve represents the case where ℓ=+1. In both (a) and (b) the dotted line shows the optical dipole potential energy without taking into account the contribution of the spin-orbit term. The insets to the figures show the potential energies of the solid curves and their projections in the focal plane. These insets should be interchanged for the dashed curves. In all plots the potential energy is in recoil energy units while the radial distances are in LG beam waist w_0 units.
|
2012.09465
|
Fig1.jpg
|
0.537845 |
6d19c5b4b60143ae834f65a3ace93ccb
|
(a) The optical dipole potential energy for right-handed circular polarization. Solid line represents the case where ℓ=+5, dashed curve is for ℓ=-5. (b) The optical dipole potential energy for left-handed circular polarization. Solid line represents the case where ℓ=-5, dashed curve is for ℓ=+5. In both (a) and (b) the dotted line shows the optical dipole potential energy without taking into account the contribution of the spin-orbit term. In both plots the potential energy is in recoil energy units while the radial distances are in LG beam waist w_0 units.
|
2012.09465
|
Fig2.jpg
|
0.498272 |
1eb48bec68b0485dba9fe6ea2581c29d
|
(a) The optical dipole potential energy for a bi-chromatic field with ℓ=1 (solid line). The plot is inverted if ℓ=-1 (dashed line). (b) Optical dipole potential energy for a bi-chromatic field with ℓ=5 (solid line). This plot is inverted if ℓ=-5. In both plots the potentials are given in recoil energy units while the radial distances are in LG beam waist w_0 units.
|
2012.09465
|
Fig3.jpg
|
0.442721 |
f26e0411027d4b1482a0e43a5e96c2d5
|
(a) The optical dipole potential energy, in the radial and axial directions, for a bi-chromatic field with ℓ=5 (left). (b) The plot is inverted for ℓ=-5 (right). The potential energy is given in recoil energy units, the radial distances are in LG beam waist w_0 units, while the axial distances are in Rayleigh range (z_R) units. The arrows in the plot (b) indicate the trapping regions.
|
2012.09465
|
Fig4.jpg
|
0.495479 |
748b4a0b327e4a6d859e8d55f1d7437a
|
2012.09467
|
1.jpg
|
|
0.434933 |
62f10df832834a91af5196337b1c9927
|
Comparison of the analytical (exact) and annealed result of Eq. (<ref>) using the 6th order Gauss-Legendre collocation method with a timestep Δ t = 0.5. The numerical result is visually indistinguishable from the analytical (exact) solution.
|
2012.09469
|
annealing_result.jpg
|
0.429897 |
af6bb72085b249cda2a0835ca8c71e76
|
Comparison of the simulated annealed, analytical, and Newton method for u(0) = 0.1. The analytical solution was obtained by using the 4th Runge-Kutta method with much smaller time step sizes. The simulated annealed solution does not match the result of the Newton method.
|
2012.09469
|
nonlinear_annealing_0.1.jpg
|
0.454794 |
2b41d3e6f4c24d68a208be3a3f23a61a
|
Comparison of the scaling of the energy gap between the ground and first excited state for the adiabatic Hamiltonian for the Crank-Nicolson scheme with changing Δ x. It can be seen that g_min is proportional to Δ x and higher time steps result in a lower energy gap.
|
2012.09469
|
scaling_crank_nicolson_dx.jpg
|
0.511214 |
dbdd7da906f44890b6a5e1761ab1563c
|
Comparison of the scaling of the energy gap between the ground and first excited state for the adiabatic Hamiltonian for the explicit Euler. For sufficient small Δ x, the expected scaling and numerical determined scaling match.
|
2012.09469
|
scaling_explicit_euler_dx.jpg
|
0.404099 |
c02ea58f53e44749ae494148d0d66c53
|
Error of the annealed Gauss-Legendre 6th order Runge-Kutta step using our variational approach, compared to a numerical evaluation of the same Runge-Kutta step. The dashed lines are obtained on a D-Wave quantum annealer, and the solid lines show an exact annealing reference solution calculated on a classical computer. The parameter c is the exponent shift of the variational approach in Eq. (<ref>). For all cases, two qubits per number register are used.
|
2012.09469
|
variational_convergence.jpg
|
0.45143 |
76c07e9a50834f3ca7e011d726f4eada
|
(Color online) Pressure as a function of baryon density ρ/ρ_0 for hadronic phase (solid line), quark phase (dash line) and mixed phase (dash-dotted line). The fitted EoSs with first-order (a) and crossover (b) phase transitions use respectively B= 150 and 90 MeV/fm^3 in the MIT model.
|
2012.09471
|
fig1.jpg
|
0.468463 |
b6db6d87f0c94daaafaabb523187fff0
|
(Color online) Maximum compression density reached with hadronic EoS (thick lines) and EoSs including first-order phase transition and crossover (thin lines) in Au + Au collisions at beam energies of 2-8 GeV/nucleon.
|
2012.09471
|
fig2.jpg
|
0.431901 |
72e92f1bdbb440ea9f93bc6dd74261b2
|
(Color online) Proton average transverse momentum ⟨ p_x⟩ as a function of rapidity in Au + Au collisions using EoSs with and without phase transition at beam energies of 2-8 GeV/nucleon. The experimental data are taken from Ref. <cit.>.
|
2012.09471
|
fig3.jpg
|
0.441927 |
14f6c0ee578c464291e8b78fdcdaa249
|
(Color online) Proton directed flow v_1 in Au + Au collisions using three kinds of EoSs at beam energies of 2-8 GeV/nucleon. The experimental data are taken from Ref. <cit.>.
|
2012.09471
|
fig4.jpg
|
0.434658 |
799c2f4c81354682a93d1ae739600160
|
(Color online) Same as Fig. <ref>, but with different phase-transition boundaries.
|
2012.09471
|
fig5.jpg
|
0.46452 |
5b559b27ce2649809c18090df3efc9b2
|
The formation of the total number of bonds 𝒩_ b with time t for different bond lengths ℓ_ b. The data is shown for n=3 tri-functional (dashed lines) and n=4 tetra-functional (solid lines) monomers. Parts (a) and (b) show the data for the FENE and harmonic bonds, respectively.
|
2012.09473
|
Figure1.jpg
|
0.532836 |
0e82cbb822524b78bfe51683e8130628
|
Number density of bonds ρ_ b (panel a) and monomers ρ_ m (panel b) as a function of bond length ℓ_ b. The data is shown for the systems after the isobaric equilibration. Open and solid symbols are for the FENE and the harmonic bonds, respectively. Lines are drawn to guide the eye.
|
2012.09473
|
Figure2.jpg
|
0.43542 |
3b7b20708601466e850311e20932f694
|
Panels (a-d) show the simulation snapshots of a 2σ thick layer along the z direction after the network cure in the canonical ensemble. The lateral dimensions of the snapshots are 67.03 σ. Snapshots are shown for different bond lengths ℓ_ b and functionalities n, as mentioned in the figure headings. Panels (a) and (c) are for ℓ_ b = 0.90σ and panels (b) and (d) are for ℓ_ b = 0.97σ. Panel (e) shows the fraction of the void volume v_ void as a function of ℓ_ b. Open and solid symbols are for the FENE and the harmonic bonds, respectively. Lines are drawn to guide the eye.
|
2012.09473
|
Figure3.jpg
|
0.476856 |
9316b3c0cf8c4b0a9c1368152ada4c77
|
The normalized thermal transport coefficient κ/κ_ linear as a function of the bond length ℓ_ b. The data is shown for different network functionality n and for the both bonds. The data is normalized with respect to the linear chain connected by the harmonic springs giving κ_ linear = 5.4±0.3 k_ B/τ m. Here the linear chain lengths are chosen as N_ℓ=50. Open and solid symbols are for the FENE and the harmonic bonds, respectively. Lines are drawn to guide the eye.
|
2012.09473
|
Figure4.jpg
|
0.434515 |
270df98126f041498b71653f06d03963
|
The normalized thermal transport coefficient κ/κ_ linear as a function of the volumetric specific heat c_ v (part a) and the sound wave velocity v (panel b). The data is shown for different functionality n, bond length ℓ_ b and for both bonds. The data is normalized with respect to the linear chain connected by the harmonic springs giving κ_ linear = 5.4±0.3 k_ B/τ m. Here the linear chain lengths are chosen as N_ℓ=50. Open and solid symbols are for the FENE and the harmonic bonds, respectively. Lines are drawn to guide the eye.
|
2012.09473
|
Figure5.jpg
|
0.536148 |
891b015e5dd54ce88f031664dc0d560a
|
Vowel-wise acoustic feature comparisons between COVID-19 negative (neg.) and COVID-19 positive (pos.) participants in form of boxplots for features with a differentiation effect r>.4 ordered from left to right according to a decreasing r, respectively. The effect size r as well as the p-value of the Mann-Whitney U difference test are given above each boxplot. p is rounded to three decimal places. r is rounded to two decimal places. Outliers (marked with red plus symbols) are defined as value that are more than 1.5 times the interquartile range away from the bottom or top of the respective box. # = number of, F0 = fundamental frequency, F1 = first vowel formant, F3 = third vowel formant, len. = length, pctlrg = percentile range, RS = rising slope, seg. = segment, ST = semitone from 27.5 Hz, SDnorm = standard deviation normalised by the arithmetic mean (coefficient of variation), slp = slope; VR = voiced regions
|
2012.09478
|
boxplots.jpg
|
0.434778 |
541175090317427499416a9170371b58
|
Dynamical formation of topological synchronized states (TSSs). (a) Schematic of the considered nonlinear system. Nonlinear oscillators are placed on a square lattice. Each oscillator is linearly coupled to other oscillators at the nearest neighbor sites. The linear coupling corresponds to the Hamiltonian of a topological insulator laser, which exhibits amplified edge oscillations. (b) TSS observed under the open boundary condition. The nonlinear oscillators near the boundary synchronize, while those away from the boundary exhibit chaotic dynamics instead of being synchronized. (c),(d) Frequency of the first component of oscillators at each site. We numerically calculate the dynamics of the first model of the TSS (Eqs. (<ref>), (<ref>)). Panels (c) and (d) are the snapshots at the time t=100 and t=200 for each. Empty sites represent the oscillators oscillating around their natural frequencies. We confirm that the frequencies of the edge oscillators are almost homogeneous and constant, which indicates their frequency-synchronization. Meanwhile, the oscillators inside exhibit inhomogeneous frequencies. One can also confirm that their frequencies vary over time, thus indicating the desynchronization and instability of the bulk oscillators. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.2. (e) Amplitudes of the oscillators. Here, we numerically simulate the dynamics of the model as in panels (c) and (d). We check that the amplitudes of most of the edge oscillators are larger than those of the bulk oscillators. The same parameters as panels (c) and (d) are used in the calculation.
|
2012.09479
|
figure1.jpg
|
0.418368 |
39e44e4776f3406e8d41bb48c9db67fe
|
On-demand pattern designing of the synchronized oscillators. We numerically demonstrate that the on-demand pattern designing of the synchronized oscillators is possible by utilizing the TSS. To arrange the synchronized oscillators in the shape of characters, UT, we damp the oscillators encircled by the green boxes and numerically calculate the dynamics of the first model of topological synchronization (cf. Eqs. (<ref>), (<ref>) and Fig. <ref>) under the periodic boundary condition. The parameters used are u=-1, b=0.3, α=0.5, α_d=-100, β=1, ω_0=0.1, and Δω=0.2. In panel (a), the frequencies of the oscillators are shown. The oscillators around the damped ones exhibit a homogeneous frequency, indicating the frequency-synchronization of those oscillators. We obtain the desired shape of the collection of the synchronized oscillators. In panel (b), the indicator of the local order parameter M(x,y) in Eq. (<ref>) at each site is shown. We also confirm the appearance of the desired synchronization pattern from this result.
|
2012.09479
|
figure10.jpg
|
0.429163 |
fc4a26f4352842a99190af33d0a264d4
|
Applications of the TSS to the defect detection. (a),(b) Schematics of detection mechanism. We assume that the dynamics of nonlinear oscillators are detected by sensors. Information of the oscillators' state from the sensors is processed to judge the synchronization of oscillators and the existence of defects. If oscillators are broken and stop their self-oscillation, the oscillators around them are synchronized as shown in panel (a). By observing such synchronization, we can detect the breakdown of oscillators. When sensors are disordered and send no signals, the oscillators remain desynchronized as shown in panel (b). Therefore, this kind of breakdown of sensors is distinguishable from the breakdown of oscillators. (c) The indicator of the local order parameter M(x,y) in Eq. (<ref>) at each site under the existence of the broken oscillators. Large values of the indicator show the synchronization of the oscillators at the corresponding site. One can judge that the oscillators surrounded by synchronized ones are broken. The parameters used are u=-1, b=0.5, α=0.5, α_d=-100, β=1, ω_0=0.1, and Δω=0.2. (d) The indicator M(x,y) at each site under the existence of the broken sensors. The indices are small everywhere, which shows that there are no synchronized oscillators, and thus no oscillators are broken. The parameters used are the same as panel (c).
|
2012.09479
|
figure11.jpg
|
0.484298 |
b2adbf3b25844666b6d573e15aed92d2
|
Proposed electrical circuit to realize topological synchronization. (a) Linear system described by the effective Hamiltonian of topological insulator laser constructed by an electrical circuit using capacitors, inductors, resistors, and negative impedance converters with current inversion (INICs). We consider two layers of square lattices, and each lattice point (a red dashed circle) has two sites (black dots). The dynamics of the voltages at the sites follows the Hamiltonian of topological insulator laser in Eq. (<ref>). The left inset represents the detail of the circuit in each layer (corresponding to the enlarged view of the sites encircled by the blue dot curve in the middle panel). The middle panel shows how to connect the sites of the different layers. The legend shows the correspondence of the figures and the circuit elements. (b) Substitution rules to construct the electrical circuit of TSS. We can construct the model of the TSS by replacing the circuit elements in panel (a). We substitute each site (a black dot) into two van-del-Pol circuits (blue dashed squares) coupled by an INIC. Green-filled boxes represent the nonlinear resistors. The other panels represent the substitution rules of capacitors, inductors, resistors, and INICs. We only use resistors and INICs to avoid frequency dependence of impedance. Red dashed squares represent the coupled van der Pol circuits in the left panel.
|
2012.09479
|
figure12.jpg
|
0.48385 |
3b2cc527a5334af1aba33a2fa08585e2
|
Dispersion relation of the lasing modes localized at the right edge. We calculate the dispersion relation of the Hamiltonian of topological lasing modes (<ref>). We consider the open (periodic) boundary condition in the x (y) direction and arrange 20 sites in the x direction. We obtain a gapless edge band with positive imaginary parts of eigenvalues. This gapless band corresponds to lasing edge modes localized at the right side. The parameters used are u=-1 and γ=0.5.
|
2012.09479
|
figure13.jpg
|
0.43492 |
9d7da78c609a425aa39ae42cd7ff1ccc
|
Cluster synchronization from a topological Hamiltonian. We calculate the dynamics of the Stuart-Landau oscillators with linear coupling described by a Hamiltonian of topological lasing modes that is different from the Hamiltonians considered in the main text (<ref>). Panels (a) and (b) show the snapshots of the frequency distributions of the first component of oscillators at the time t=1000 and t = 2000 for each. The bulk and left-side oscillators exhibit homogeneous and constant frequency, which is different from that of the right-side oscillators. Therefore, the bulk oscillators form a synchronized cluster, indicating the absence of the TSS. The parameters used are u=-1, γ=0.5, α=0.5, β=1, ω_0=1, and Δω=0.2.
|
2012.09479
|
figure14.jpg
|
0.412836 |
8e06ea0dbbc7499c9df52a314b296428
|
Coexistence of synchronized edge oscillators and damped bulk oscillators. (a) Frequency of the first component of oscillators at each site in a damping parameter region. We numerically simulate the first model of the TSS (Eqs. (<ref>), (<ref>)) at the parameters u=-1, b=0.5, α=-0.2, β=1, ω_0=1, and Δω=0. The figure shows the snapshot at time t=100. The edge oscillators exhibit homogeneous frequencies and thus synchronize. (b) Amplitudes of the first components of oscillators at each site at time t=100. We can confirm that the bulk oscillators exhibit almost zero amplitudes. The parameters used are the same as in panel (a).
|
2012.09479
|
figure15.jpg
|
0.433622 |
9eb08df2845f4dd6af510e654fbbd246
|
TSS without fluctuations of the natural frequencies. We simulate the first model of the TSS (Eqs. (<ref>), (<ref>)) under the condition of the homogeneous natural frequencies. Panels (a) and (b) show the snapshots of the frequency distributions of the first component of oscillators at the time t=100 and t=200 for each. As in the case of the fluctuating natural frequencies (Fig. <ref>(c),(d)), the edge oscillators exhibit homogeneous and constant frequencies, while the bulk ones vibrate at space- and time-varying frequencies. Therefore, the edge oscillators synchronize, while the bulk ones desynchronize, which indicates the emergence of the TSS. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.
|
2012.09479
|
figure16.jpg
|
0.523824 |
bfc96dbb14e34a64b1ae46240d377ed2
|
Dispersion relation of the linearized equation of one-dimensional chain model. We numerically calculate the dispersion relation of the coefficient matrix obtained from the linearization of the one-dimensional chain model around the stationary solution (Eqs. (<ref>), (<ref>)). The upper (lower) panel shows the real (imaginary) part of the eigenvalues. All the eigenvalues have nonpositive imaginary parts, which implies the linear stability of the stationary solution. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=0, and Δω=0.
|
2012.09479
|
figure17.jpg
|
0.400378 |
c0abf1318b6c425e8c3c79a0da4b72bf
|
Robustness of TSS against disorders in linear couplings. We numerically simulate the model of the TSS (Eqs. (<ref>), (<ref>)) with the inhomogeneous linear couplings. Panels (a) and (b) show the frequency distributions at time t=100 and t=200 for each. The edge oscillators exhibit homogeneous and constant frequencies, while the bulk ones oscillate at time- and space-varying frequencies. Therefore, we obtain the TSS robustly against the noise in linear couplings. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.2.
|
2012.09479
|
figure18.jpg
|
0.41893 |
20a68d9534c84bf6b3135232ed98c7aa
|
Dispersion relation of the effective Hamiltonian. We calculate the dispersion relation of the effective Hamiltonian obtained from the linearization of the first model (cf. Eqs. (<ref>), (<ref>)). We consider the open (periodic) boundary condition in the x (y) direction and arrange 20 sites in the x direction. While the bulk bands are gapless as opposed to conventional topological systems, the edge localized modes still exist in this dispersion relation. The edge modes exhibit negative imaginary parts of the eigenvalues. In contrast, some bulk modes possess positive imaginary parts of the eigenvalues depicted by the red dashed circle, which indicates instability of the bulk oscillators. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.
|
2012.09479
|
figure19.jpg
|
0.428422 |
56a8574d23c944a3b2f0273bf4e63efd
|
Phase diagram of the model of the TSS. The ratio of frequency fluctuations of the edge oscillators to those of the bulk ones is depicted at each parameter. We calculate the dynamics of the model of the TSS (Eqs. (<ref>), (<ref>)) under different strengths of the linear coupling and the inhomogeneity of the natural frequencies. The red region (right bottom) shows parameter regimes where the TSS emerges. The circles and squares represent the classification of data of frequency fluctuations of the edge and bulk oscillators obtained by a numerical clustering method. The circles correspond to the parameters for which the presence of the TSS is identified. The phase boundary obtained here corresponds to the parameter points at which the bandgap of the effective Hamiltonian closes and thus the bulk band can change its topology.
|
2012.09479
|
figure2.jpg
|
0.445635 |
22ff9df41b854dbbb395b5a61973e7f6
|
Amplitude distribution indicating amplitude chaos. The amplitudes of oscillators at the first components of lattice points are shown. To obtain the amplitude distribution, we numerically calculate the dynamics of the first model of TSS (cf. Eqs. (<ref>), (<ref>)). The red-filled circles represent the site with small amplitudes. The existence of such small-amplitude oscillators indicates amplitude chaos. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.2.
|
2012.09479
|
figure20.jpg
|
0.48224 |
81abffe99d6d4bf4836010abd8472d9e
|
Inverse participation ratios of the Lyapunov vectors in the wavenumber space. We numerically calculate and plot the inverse participation ratios of the Fourier components of the Lyapunov vectors obtained from the model of the TSS (Eqs. (<ref>), (<ref>)). The relative indices correspond to those in Fig. <ref>. We confirm the large IPRs at small relative indices (smaller than about 0.05) and large relative indices (larger than about 0.95). The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=0.2, and Δω=0.2.
|
2012.09479
|
figure21.jpg
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.