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
0.378794
d26d79b11ed3443ea6a39aaf74a6cfd0
The first two panels from the left illustrate the different formation paths described in section [sec:formation_channels]3.1. In the direct physical collision scenario, the dominant process by which the principal object increases in mass is through physical collisions, as is shown by the initial binary-binary interaction that leads to the merger of two stars. In contrast, in the middle, for the formation path labeled binary coalescence, the main increase in mass comes from binary coalescence.The right-most column illustrates the collision history of the most massive IMBH formed in our simulations.
2012.10497
tree.jpg
0.464724
679aec000b89491d89931845bdb479ab
Specific heat capacity of NdCoGe_3 for (a) magnetic field within the basal plane and (b) field applied along the c axis.
2012.10499
Cp1.jpg
0.543801
d7f70238ad7d4e4b8ab2a685ff2931d9
Zero-field data revealing magnetic phase transitions at T_N1 = 3.70 K and T_N2 = 3.50 K in NdCoGe_3 indicated by the vertical dashed lines. (a) Specific heat capacity and (b,c,d) the in-plane ac magnetic susceptibility measured in zero applied DC field. The ac data are shown as the (b) in-phase contribution χ' with (c) displaying the derivative of χ', and (d) contains the out-of-phase contribution χ”.
2012.10499
Figure1.jpg
0.497287
cd469f8ad55240808cba91c5442c6d1e
(a,c,e) Neutron diffraction intensity as a function of h, k or l at various temperatures and (b,d,f) the fitted position of h, k, and l as a function of temperature. The red lines in (a,c,e) are the corresponding Gaussian fit(s) and the fitted error bars in (b,d,f) are smaller than the size of the symbols. In (a), the small shoulders for smaller h are artifacts from the sample analyzer collimator.
2012.10499
Figure3.jpg
0.497347
8ae2fc5a5ea54ee08f6e69c1a84ad08c
Temperature-dependent magnetization in zero-field-cooled warming (red) and field-cooled cooling (blue) measurements with an applied field (a) H ⊥ c and (b) H||c. The inset in (a) shows the magnetization to T = 0.4 K using the same units as the main panel.
2012.10499
Figure4.jpg
0.415548
3dd2f767d1db43b78e4e48ac8e416a97
Isothermal magnetization M(H) and derivative of isothermal magnetization dM/dH of NdCoGe_3 with an applied field H⊥c. (a) M(H) (left axis) and dM/dH (right axis) at T = 0.4 K (b) The in-plane ac magnetic susceptibility χ' as a function of applied DC field at T = 0.4 K (c) dM/dH from 1.8 K to 2.8 K with a temperature step of Δ T = 0.04 K. (d) dM/dH at T = 3 K.
2012.10499
Figure5.jpg
0.473994
288057c4e3104d7bb4181e037434fdfc
(a,b) Isothermal magnetization data (left axis) and derivative dM/dH (right axis) on increasing field (red line) and decreasing field (blue line) of NdCoGe_3 with an applied field H⊥c.
2012.10499
Figure6.jpg
0.421257
f2fadf026cc34d79aa26a9da0edf7240
Neutron powder diffraction pattern of NdCoGe_3 at T = 1.8 K < T_N2 . Inset: Data for a limited Q range revealing magnetic reflections at 1.8 K (bottom) through a comparison to the pattern obtained at 40 K (top) for which the Bragg peaks are entirely from the crystal lattice.
2012.10499
NeutronPowder.jpg
0.410236
387e066b43d04a809a4d89a19710903c
Magnetic phase diagrams of NdCoGe_3 as inferred from anomalies in the specific heat, derivatives of temperature-dependent and isothermal magnetization as indicated in the legends for (a) H⊥c and (b) H||c. The black dotted line in (a) indicates the presence of a first order phase boundary based upon observations of thermal or field hysteresis.
2012.10499
Phase.jpg
0.445534
e726616ccac54a5a9af51274485f77cd
Neutron single crystal diffraction data for NdCoGe_3 from the WAND^2. A section of the reciprocal h0l plane between 00l and -10l with the magnetic reflections close to the -0.5 0 l position. The color scale is adjusted to show both the weak magnetic intensity and the stronger nuclear scattering. The rings originate from the Al-sample holder. Purple arrows indicate the propagation vector from the crystallographic -103 and 004 reflections.
2012.10499
Wand2_new.jpg
0.449911
13622d1a9f814824a85273c24bc5575f
Candidate magnetic structures resulting from (a) 1k⃗, (b) 2k⃗, (c) 3k⃗ and (d) 4k⃗ models. The vector moments (blue arrows) on Nd atoms (blue spheres) are shown for a segment of the magnetic unit cell; the nuclear (paramagnetic) unit cell is outlined in grey, the Co atoms are red and Ge atoms are grey. The various models describe the neutron powder diffraction data at 1.8 K equally well.
2012.10499
mag_str.jpg
0.413801
95c2b092a1e64911a5efcba08ff64bfa
Location of the FASERν detector and event topology. Top: The FASER experiment is placed about 500 m downstream of the ATLAS interaction point in the previously unused side tunnel TI12, which connects the SPS with the LHC tunnel. Center: The detector is centered around the beam collision axis where the neutrino flux is maximal. It consists of the FASERν emulsion neutrino detector, followed by a magnetized spectrometer and a calorimeter. Bottom: The emulsion detector consists of tungsten plates interleaved with nuclear emulsion films. Both interactions of neutrinos and neutral hadrons lead to the appearance of a neutral vertex at which several charged particles emerge. Different types of events can be distinguished based on the event topology, as explained in the text.
2012.10500
Figure.jpg
0.517272
776b9d59f2344eb09fbc35259b8c168f
Comparison of 3 observables for NH interactions generated using , , and and NC interactions using , and . Top: NH interactions at 60, 300. Middle: Different NHs (mesons and baryons) generated via at 300. Bottom: NC interactions at 100, 1 generated via , and . See text for more details.
2012.10500
appendix_fig.jpg
0.429211
c674bcc9d693488ebece1fa7a7245cc5
Left: Correlation matrix showing the linear relationship between observables presented in <ref>, the incoming particle energy (E_beam) and the event type (1 for NC neutrino interactions, 0 for neutral hadrons). Right: Permutation feature importance (the normalized mean score decrease for each of the observables) for the signal identification classifier (blue) and neutrino energy estimator (red) network. We use accuracy as the score metric for the classifier network and mean average percents error for the estimator. Scores decreases are normalized so that they sum to 1.
2012.10500
corr_perm_importance.jpg
0.385364
be11346bd4b049aab3847874a59d5d15
Left: Signal selection efficiencies as a function of beam energy. Each line indicates the fraction of events passing the following criteria sequentially: i) neutral vertex identification (blue) requiring ≥ 5 charged tracks, ii) lepton veto (green) requiring no electron candidate and no non-interacting charged track, and iii) signal identification (red) as performed by the NN classifier. The dashed black line shows the combined efficiency. Right: The energy spectra of particles interacting within the FASERν detector. We show the expected numbers of neutral hadron interactions (green) and NC neutrino interactions (red) with the FASERν detector during LHC Run 3 as dashed lines. The solid lines show the spectra for events passing the signal selection (including neutral vertex identification, lepton veto and signal identification). The uncertainty associated to the background generation is shown as a shaded band.
2012.10500
efficiency_intrct_spectrum.jpg
0.393763
063b20839c7140338a9c2fcb9587e72d
Left: Neutrino energy reconstruction for NC neutrino interaction events obtained by a neural network-based multivariate analysis using the observables defined in <ref>. Right: Relative RMS energy resolution using the neural network-based multivariate analysis (solid) and only the hadronic energy of the events (dashed).
2012.10500
energy2D_resolution.jpg
0.458506
7257e621190c4eecbeb1c634f4685840
Expected energy spectrum of neutral hadrons interacting within the FASERν detector during LHC Run 3 with 150 of luminosity.
2012.10500
hadspec_072720_neutral.jpg
0.495289
e46949d501bf4c59b2511b6af8e1d293
Normalized kinematic distributions for the observables defined in <ref>. The dashed lines show the distributions obtained with for the NC neutrino interaction signal at incoming neutrino energies of E_ν=100 (blue) and E_ν=1 (red). The solid green lines correspond to the distributions for the neutral hadron interactions simulated with for the expected energy spectrum presented in <ref>. The shaded region shows the range of predictions for the background distributions obtained from different generators: , and .
2012.10500
obs.jpg
0.442296
e36519aa97ee4818a90ee63d7d142c74
Left: Stacked histogram of events passing the event signal selection described in <ref> as function of the reconstructed energy for LHC Run 3 with 150 integrated luminosity. The red and green shaded regions show the NC neutrino interaction signal and the neutral hadron interaction background, respectively. The hatched region indicates the uncertainty arising from the simulation of neutral hadron interactions, corresponding to the range of predictions obtained by three different generators. Right: FASERν's estimated neutrino-tungsten NC cross section sensitivity. Existing constraints are shown in gray. The black dashed curve is the theoretical prediction for the DIS cross section, averaged over neutrinos and anti-neutrinos, per tungsten nucleus. The inner red error bars correspond to statistical uncertainties, the blue error bars additional take into account uncertainties associated with the simulation of the background, and the outer green error bars show the combined uncertainties with the neutrino production rate (which corresponds to the range of predictions obtained from different MC generators as obtained in Ref. <cit.>).
2012.10500
signal_spectrum_crosssection.jpg
0.483784
c2cc8de2259e45aca2831214faf5470c
Spectra of radiation from sidewalls of the accretion columns, calculated simultaneously with velocity and temperature profiles plotted in Figs <ref> and <ref>. The results correspond to different sets of the model parameters: `1' (solid line), `2' (dashed), `3' (dot-dashed), `4' (dot-double-dashed), `5' (double-dot-dashed), `6' (dotted).
2012.10501
figsp1.jpg
0.455476
676006c692dd46b295d0a3b315e4fade
(a) Electron data (in blue) from left to right: SUMER on-disk “correlation heights” in open funnels <cit.>, SUMER off-limb temperatures from (; dark blue) and (; light blue), UVCS <cit.>, PSP <cit.>, and Helios/Ulysses <cit.>. Proton data (in red) from left to right: UVCS <cit.>, PSP (dark red), Helios (light red), Ulysses (dark red), and Voyager (light red) high-speed wind data. Oxygen-ion data (in green) from left to right: SUMER off-limb kinetic temperatures <cit.>, UVCS <cit.>, and Wind <cit.>. (b) From left to right, O^+5 and proton anisotropy data from UVCS <cit.>, PSP data with v ≥ 450 km s^-1 <cit.>, Helios data with v ≥ 600 km s^-1 <cit.>, and Ulysses data with v ≥ 600 km s^-1 <cit.>. All black curves are solely to guide the eye.
2012.10509
cranmer_temps2020_fig1.jpg
0.459751
0a57cfccc5fd4e298dcbd19763830963
Comparison of the HL (purple) and HM (dark blue) methods to the modified LCHO-CI method for evaluating J. The system is a double QD in Si/SiO_2 described by a quartic potential, where l_0= 6 nm. a) Probability densities of the ground single-electron orbital states for the HL/HM and LCHO-CI calculations at different dot separations. At small dot separations, the HL/HM states overestimate the separation of the two localized wave functions compared to the numerically calculated orbitals. As dot separation increases, the approximate HL/HM states converge to the numerical results. b) J versus the interdot separation 2d is plotted. Three different LCHO-CI calculations are done for N=2 (light blue), N=4 (green), and N=10 (yellow). For all LCHO-CI calculations, M=15^2.
2012.10512
F_HL_HM_comp.jpg
0.430391
49ce28a13a74468ba463c69ddd820f0e
Convergence of J with respect to the number of single-electron orbitals N when M=16^2. The device parameters are V_p1= V_p2= 0.15 V, V_t= 0.09455 V, D_x= D_y= 40 nm, D_t= 20 nm, and T= 15 nm. Inset: The convergence of J with respect to M when N=18 (star in the main figure).
2012.10512
F_Jconvergance.jpg
0.403861
f490e4da394046ef9c9b58455438b9b3
Using harmonic orbitals (HO) to approximate single-electron states. a) The operator  transforms between the HO basis {|ϕ_i⟩} and the approximate single-electron orbitals |ξ_j'⟩. The first three single-electron orbitals for a quartic potential are shown at left, and the eight lowest energy HO states are shown at right. b) Convergence of the first twelve approximate single-electron energies ϵ_j' versus the number of HOs, M, used to compose the basis {|ϕ_i^ω⟩}. Here, M = M_xM_y and M_x=M_y are the numbers of 1-dimensional HOs taken along the x and y axes, respectively, to construct the 2-dimensional HOs.
2012.10512
F_SEconvergance.jpg
0.532902
ddd86be3f3c245ae97598d4f37f3ef30
Relationship between bias voltage V_ bias and inter-dot detuning ϵ for different dot sizes. The calculated data points are indicated by circles and the solid line is a fit to the equation ϵ = α V_ bias where α is the lever arm. The region in the dashed black box is enlarged and shown in the inset to demonstrate the linearity of the fitted data.
2012.10512
F_bias_fit.jpg
0.492798
ae5b43f664c747ec88c74b501594b8ad
Schematic of a double QD device formed by plunger, tunnel, and screening gates. The geometric parameters are D_x= D_y= 40 nm, D_t= 20 nm, and T= 15 nm. Applied voltages in (a) and (b) are V_p1= V_p2= 0.150 V, and V_ tun= 0.09455 V. a) 3D device model, where the plunger, tunnel and screening gates are shaded differently for contrast, along with a semi-transparent SiO_2 layer. A 2D electrostatic potential obtained by the self-consistent 3D Poisson calculation is plotted beneath. b) The 2D potential overlaid with an outline of the gate structure, showing the heads of the plunger and tunnel gates as well as the screening gate. The gate lengths (extending beyond the screening gate) are denoted D_y, and the widths of the plunger and tunnel gates are D_x and D_t, respectively. The 2D potential is taken 1 nm below the Si/SiO_2 interface. c) Side profile of the gate structure taken along the dashed line in b). The SiO_2 layer thickness is labelled T. The QD (yellow ellipse) is formed underneath the head of the plunger gate.
2012.10512
F_deviceSchematic.jpg
0.453271
2977cd872fdc484f90d5fe0bb1efbb4d
Dependence of J on tunnel gate voltage as the dot size D=D_x=D_y is varied. a) J versus V_ tun, where the latter is varied from 0.03-0.150 V. The fixed device parameters are V_p=0.150 V, V_ bias=0 V, D_t= 20 nm, and T= 15 nm. The dashed black line indicates J= 1 μeV. b) Derivative of J with respect to V_ bias. For each device geometry, V_ tun is tuned so that J= 1 μeV at V_ bias=0. c) Derivative of J with respect to ϵ, where ϵ = α V_ bias and α is the lever arm between the applied bias voltage and the effective inter-dot detuning. The upper-right table provides a legend and includes the calculated charging energy U and lever arm for each geometry.
2012.10512
F_dot_size.jpg
0.431734
7b42e4e164de46e9b7f06de1334bd47d
Dependence of J on the plunger gate eccentricity, D_y/D_x. For D_y/D_x> 1.0, D_x= 40 nm, while for D_y/D_x< 1.0, D_y= 40 nm. a) J versus V_ tun, as the latter is varied from 0.03-0.150 V. The fixed device parameters are V_p=0.150 V, V_ bias=0 V, D_t= 20 nm, and T= 15 nm. b) Derivative of J with respect to V_ bias. Each device has V_ tun tuned so that J= 1 μeV at V_ bias=0, as indicated by the dashed black line in panel (a). c) Derivative of J with respect to ϵ, where ϵ = α V_ bias and α is the lever arm between the applied bias voltage and the effective inter-dot detuning. The upper-right table provides a legend and includes the charging energy and lever arm versus dot eccentricity.
2012.10512
F_ecc.jpg
0.45815
5ec99ebfd071406a86b4d5efcdcccca3
Dependence of the optimization function f_min(ω) on ω and the size of the harmonic orbital basis (M_xM_y=M). N'=6 for a quartic potential with parameters m^* = 0.191m_0, ħω_0= 0.375 meV, and d = 50 nm. Color indicates a different number of HO basis states ranging from M_x=M_y=1 (purple) to M_x=M_y=16 (yellow).
2012.10512
F_opt_omega.jpg
0.489292
29d5eb415a9e4689aed3f3ecc7810fd4
Dependence of J on varying oxide thickness T. a) J versus V_ tun, as the latter is varied from 0.0-0.100 V. The fixed device parameters are V_p=0.100 V, V_ bias=0 V, D_x= D_y= 40 nm, and D_t= 20 nm. b) Derivative of J with respect to V_ bias. Each device has V_ tun tuned so that J= 1 μeV at V_ bias=0, as indicated by the dashed black line in panel (a). c) Derivative of J with respect to ϵ, where ϵ = α V_ bias and α is the lever arm between the applied bias voltage and the effective inter-dot detuning. The upper-right table provides a legend and includes the charging energies and lever arms versus T.
2012.10512
F_oxide.jpg
0.446911
62b0d1dcc8d1495b9f8f4151f45f35ce
Dependence of J on a varying tunnel gate width D_t. a) J versus V_ tun, as the latter is varied from 0.03-0.150 V. The fixed device parameters are V_p=0.150 V, V_ bias=0 V, D_x= D_y= 40 nm, and T= 15 nm. The dashed black line indicates J= 1 μeV. b) Derivative of J with respect to V_ bias. For each device geometry, V_ tun is tuned so that J= 1 μeV at V_ bias=0. c) Derivative of J with respect to ϵ, where ϵ = α V_ bias and α is the gate lever arm. The upper-right table provides a legend and includes the charging energy U and lever arm α versus D_t.
2012.10512
F_tun_gate.jpg
0.464564
75f09d3902b8461088a2384535482996
1D line cuts of the 2D potentials that give J= 1 μeV at V_ bias= 0 V. Unless varied, the default device parameters are D_x=D_y= 40 nm, D_t= 20 nm, and T= 15 nm. For D_y/D_x>1.0, D_x= 40 nm, and for D_y/D_x<1.0, D_y= 40 nm. The corresponding gate voltages are given in Table <ref>. The 1D slices are taken along the x-axis and pass through the lowest potential minima in the 2D potentials.
2012.10512
F_zero_bias_pots.jpg
0.492839
1e35ce088fef4839a7efc94280d4c3e7
Consort diagram for the studies reviewed
2012.10517
DTIPRISMAv2.jpg
0.503996
b5c957da6ea64a0f845396a51f6efd9c
Our model predicts a 3D Gaussian for each keypoint parameterized by μ, Σ. To supervise this prediction using 2D annotations, we differentiably project the 3D distribution onto each of the C camera planes using para-perspective projection, yielding C Gaussians with parameters {(μ_c, Σ_c)}. We then maximize the likelihood of the annotated keypoints in each image under the projected Gaussians. move images beside above the cameras, this image needs to be much smaller (half its current height) update notation
2012.10518
2d_proj.jpg
0.525169
d75e4300c2834995b718e7115bb9bb13
Model architecture – We predict keypoints as probability distributions in 3D parameterized by a mean μ and covariance Σ representing the position and spatial uncertainty of a keypoint. As input, our model accepts multiple images of a subject taken from multiple (known) camera views. We feed these images through a 2D backbone using the architecture from <cit.> which predicts one heatmap per keypoint within each view. To predict μ, we triangulate the spatial mean of each heatmap using known camera parameters. To predict Σ, we aggregate the bottleneck features of the backbone network, and use a fully connected network to predict a decomposition of Σ that is guaranteed to be positive definite.
2012.10518
architecture.jpg
0.466384
b64d68fc1b1d4d0fa2c0c887b82f19d3
(a): Schematic of arrival times of excitation PL pulse and STED pulse, with delay τ_s between them. (b): Time-resolved ZPL-PL of defect-A with positive (red) and negative (black) τ_s. The STED pulse switches off the emission (c): Time resolved PL of defect-A at five different values of τ_s, showing rapid depletion at arrival time of STED pulse. (d): Plot of time-averaged ZPL-PL intensity with varying τ_s. Moving the STED pulse through the PL pulse results in significant quenching of the PL signal that recovers as τ_s is increased. (e): Sweep of STED ratio with varying time-averaged power of the STED laser.
2012.10520
Fig2.jpg
0.452006
9e5eb6a4ae614bb89682e3f10959f1e4
(a): Calculated optical phonon dispersion for bulk hBN with LO(E_1u) mode highlighted. Taken from Serrano et al. <cit.> (b): Comparison of PL(grey), PLE (blue) and STED (red) OPSB spectra of defect-A, which has a ZPL energy of 2.170 eV. Red (blue) dashed line shows position of the LO(E_1u, Γ) transition in STED (PLE).
2012.10520
Fig3.jpg
0.443387
3b16f414795c4bf9bf427e11c9c83eb7
(a)-(d): Comparison of PL, PLE and STED spectra from four similar defects with ZPL energies of (a) 2.166 eV (b) 2.175eV (c) 2.142 eV and (d) 2.171eV. Each defect shows shift in 200meV peak between absorption and emission. In (b) the peaks marked with ∗ are from another nearby defect or defects.
2012.10520
Fig4.jpg
0.420593
c8fdfd67d32c42d88b338e23980ecd06
Level schematic of three-level system with rate parameters labeled.
2012.10520
SI-fig1.jpg
0.46498
098561ba5b57426fa3a29063a3269a60
(a): Franck-Condon energy diagram comparing PL, PLE, and STED techniques. In PL, the non-resonant excitation (green) populates the excited state and emission from the ZPL (orange) and PSB (red) is collected. In PLE, the excitation probes the excited vibronic states and the ZPL is collected. In STED, the non-resonant excitation is again used, and the STED pulses (red) deplete the excited state through stimulated emission into the ground vibronic states, reducing the ZPL intensity. Grey dashed arrows show fast relaxation from higher vibronic states. (b): Diagram of the experimental set-up. (c): Representative PL (grey), PLE (blue) and STED (red) spectrum from defect-A with ZPL at ∼2.17eV, emission OPSB between 1.97 and 2.03eV and absorption OSPB between 2.32 and 2.38eV.
2012.10520
fig1.jpg
0.499467
7d75c3ab28a24c3ca80e2f6821110582
Example of S for 𝔽_2=⟨ a,b,a^-1,b^-1⟩
2012.10522
freegroup.jpg
0.438251
1364e6d7c42e44ccaa084c8372cd6ae4
Shift on 2^
2012.10522
shiftmap.jpg
0.455737
e6e25cce207a41c48f5ddbe1d48d7f10
▹_s^4(x) with examples of S_x and S_1 x circled
2012.10522
tiles.jpg
0.430229
adadb959b61247bba86bfa8dd6373dc9
Illustration of the proofs of (a) Observation <ref>; (b) Theorem <ref>.
2012.10525
Caterpillar-2.jpg
0.448253
cceecba803d14c2ba3e1821bcbf0d11d
A schematic illustration of a embedding of T on S.
2012.10525
nice.jpg
0.469806
45d7ddcd0b004dd9aaab0b582b1b9f88
An illustration of the tree T in the proof of Theorem <ref>.
2012.10525
np-fixed-tree.jpg
0.444408
4fd2fcc04412445fb42707740b414aec
(a) A partial embedding Γ_i+1 of Γ that is not , while Γ_i is. (b) A modification Γ' of Γ such that Γ'_i+1 is .
2012.10525
np-tree-final-v3.jpg
0.529461
8b9edd93cd274e7e82fd9c9c58da870d
Actual basketball game trajectories. The trajectory data is from a Toronto Raptors NBA game in 2015, which contains offense (red), defense (blue), and ball (green) 2D overhead-view trajectories.
2012.10531
basketball_trajectory.jpg
0.476134
02d5ec800e70479b93c0a70e0e15c491
An overview of the (3D) ITN and STN architectures. Note, 3×3×3 conv. layers (red and green arrows) are followed by a ReLU activation, except for in the final layer of the STN. Outputs of the STN, v and T, are passed to the Transformation Computation Module in Fig. <ref>.
2012.10533
architectures_updated.jpg
0.389598
6c768e0fe1ab4b4bbed733cfe9a9f0fe
An axial slice through the initial (top row) and final (bottom row) atlas image (first column) and the 6 channels of the atlas labelmap produced when training Atlas-ISTN on multi-label CCTA data.
2012.10533
atlas_2d.jpg
0.449408
ce013004b84240ac91f9e1f88d576099
Scatter plots showing HD (left), ASD (middle) and DSC (right) results of Atlas-ISTN on the 1000 case test set, comparing the LVM label from the ITN (x-axis) versus refinement (y-axis). The green/red gradients indicate increase/decrease in performance with refinement. HD and ASD of the ITN predictions almost always improve with refinement. Degradation observed for some cases in terms of DSC is always small, whereas improvements in DSC can be significant.
2012.10533
ccta_scatter_plots_horizontal.jpg
0.421059
9e492f818800487fb606a7e4c6560c9d
An overview of the Atlas-ISTN framework in training (top) and test time refinement (bottom) illustrated with CCTA image data. During training, both ITN and STN weights are optimized by leveraging ground-truth labelmaps for each subject, while the atlas is updated at the end of each epoch. A symmetric loss is used to register the atlas to the subject and vice versa. At test time, instance-specific optimization (indicated by *) leverages the ITN prediction as a target labelmap to update the STN weights, providing a refined transformation from the atlas to a given subject. Spurious segmentations in the ITN prediction can be circumvented via registration of the atlas to the subject. References to related sections and figures are provided in grey text. Red boxes: atlas labelmap (deformed and undeformed). Blue boxes: input image or corresponding GT labelmap (deformed and undeformed). Green boxes: ITN logits (deformed and undeformed).
2012.10533
full_framework_updated.jpg
0.442538
2e6d0471a6e64f6caa7e16a779d4a0d9
Axial (a) and sagittal (b) planes showing examples where refinement corrects for holes (false negatives) in the LVM channel predicted by the ITN, for the Atlas-ISTN trained with an additional 2000 cases. Red contour: LVM label of deformed atlas labelmap after refinement. Yellow contour: ITN prediction of the LVM channel. Heatmap: ITN logits of LVM channel. In both (a) and (b), the ITN prediction contour consists of only a single connected component of the LVM.
2012.10533
lv_hole_segmentation.jpg
0.383227
377369b1433c4925b12f5113e64e89ed
Test cases in order of increasing difficulty (top to bottom), displaying an axial slice from a CCTA image. In columns 3, 5 and 7, manual and predicted contours of the LVM are shown in green and red, respectively, while white contours are used for other predicted structures for clarity of the LVM comparison. Column descriptions are provided above. White arrows highlight false positive and false negative LVM predictions from the ITN. Orange arrows highlight errors between the 1-pass deformed LVM and the manual LVM contours. Rows 1-2 contain examples for which ITN, 1-pass and refinement predictions all perform well, representative of a significant proportion of test cases. Rows 3-4 show cases where the ITN produces spurious segmentations, but refinement is able to circumvent this. In row 4 however, the 1-pass LVM contour is visibly offset, while the refinement produces a better fit. Rows 5-6 show challenging cases, where row 5 contains an example where the heart is particularly small in the field-of-view. On top of this, the ITN predicts a large spurious segmentation extending from the base of the LVM. The deformed contours after 1-pass do not fit the target structures accurately, but after refinement the LVM and other chambers are well positioned. Row 6 shows a case which required a significant rotation and translation of the atlas, so much so that in column 2 a different axial slice is shown to include the undeformed atlas. Additionally the LVM wall is particularly thin, where the ITN predicts a hole near the apex, and an over-segmention in the basal septal region. The significant transformation proves challenging for a single pass of the model, resulting in poor alignment of the atlas to the target SoI. The deformed contours after refinement however align well with the target SoI and bridges the ITN's predicted hole in the LVM (albeit not perfectly fitting to the manual LVM contour). Notice for cases which require larger global transformations (rows 3-6), the deformed grid after refinement contains a more noticeable affine transformation, and the non-rigid component appears to deform the grid less compared to the 1-pass.
2012.10533
seg_results_all.jpg
0.38002
de9994939a134386b35ce9dc93ed4234
Qualitative results for the brain structure segmentation for fives cases from the five different datasets (one case per row) comparing U-net, VML (i.e. a VoxelMorph-like STN-only baseline), initial atlas alignment `Id', and the three outputs ITN, `1-pass' and `Refine' of our Atlas-ISTN approach. The reference segmentation is displayed with green contours and the predicted segmentation boundaries with red contours. Note how the voxel-wise segmentation methods, U-net and ITN, tend to merge neighboring structures, while the VML baseline produces less accurate atlas alignment than `1-pass' or `Refine'. The test time refinement of the Atlas-ISTN seems to produce the visually best results, followed by the `1-pass' prediction which acts as an initialization for the refinement.
2012.10533
segmentation_brain_mri.jpg
0.457451
9a5c50eeb9874d199c03a4fb0628baf2
Axial (a) and sagittal (b) planes showing examples where refinement corrects for (a) holes (false negatives) and (b) spurious segmentations (false positives) in the LVM channel predicted by the ITN. Red contour: LVM label of deformed atlas labelmap after refinement. Yellow contour: ITN prediction of the LVM channel. Heatmap: ITN logits of LVM channel. In both (a) and (b), the ITN prediction contour consists of only a single connected component of the LVM.
2012.10533
spurious_and_hole_segmentation_landscape_2.jpg
0.369804
45bb61d8803d48de822a7a9eb3250772
The left-most column illustrates the base letter B (top), the initial atlas labelmap at the start of training defined in Eq. (<ref>) at t=0 (middle), and the recovered letter B atlas labelmap after training (bottom). The other images are examples from the training set (top row) and clean (middle row) and corrupt test set (bottom row) generated randomly from the base letter B.
2012.10533
synth2d-overview.jpg
0.379657
eb3edc2adffd4ef7be5e4d60fcc7f30f
Qualitative results for the 2D toy data with test data coming from the same distribution as the training data. Both the ITN and 1-pass Atlas-ISTN yield accurate segmentations. test time refinement with increasing regularization weight λ affects the smoothness of the final transformation.
2012.10533
synth2d-results-clean.jpg
0.407635
899618f61cf64fe7be42155b4cc30c69
Qualitative results for the 2D toy data with corrupted, out-of-distribution test data. The ITN yields many false positives and false negatives and is topologically implausible. The 1-pass Atlas-ISTN yields a reasonable atlas alignment despite the corrupted input data. test time refinement with increasing regularization weight λ can yield accurate and topologically plausible segmentations.
2012.10533
synth2d-results-noisy.jpg
0.507786
68116ed96b594e9599b2aa24c916924a
The Transformation Computation Module. The predicted SVF (v) is integrated via scaling and squaring, before linear upsampling. The resulting transformation (ϕ) is composed with the predicted affine parameters (T_i). Computation of the inverse and forward transformations are shown at the top and bottom, respectively. Compositions are defined in Eqs. (<ref>) and (<ref>).
2012.10533
transformation_computation_module.jpg
0.423184
9ee07ca589e441da8555172040083ad3
Intensity projections of images (top) and labelmaps (bottom) of a CCTA training case (left) and the constructed atlas (right). Structures depicted include the left ventricle myocardium (red), right ventricle blood pool (blue), right atrial blood pool (yellow) and left atrial blood pool (green).
2012.10533
volumes.jpg
0.507564
dbd90be8cd7d4f14b66b397151af2818
User_x and User_y is in the same GPS co-ordinate block. They connect with the same wireless access point which provides them C^τ_k_gps and K^τ_k_net. User_x generates a random number Rand_x^τ_k and User_y generates a random number Rand_y^τ_k. Both users concatenate the parameters sent by the server with the network address of the access point and pass them to a hash function to generate a hashed token. Both users upload their generated random numbers and hashed tokens to the server. The server stores the hashed tokens and the random values sent by the users in a database to later use them for contact detection.
2012.10534
Sys_design.jpg
0.444102
753ede67009541f79ce3e09142a520ce
Workflow of the system once an user reports to be positively diagnosed and chooses to share the hashed tokens and pseudo-random numbers generated during the contact detection period of the system.
2012.10534
positive_diagnosis.jpg
0.447425
b8242e463e8e4fbb92c83ec4ed1b432c
Evolution of mean density (upper plot), mass (middle plot), and mean temperature (lower plot) of gravitationally instable regions.
2012.10535
collapsed_regions.jpg
0.513371
70d9203490f247cd9bce5dc3ea777f7a
Numerically obtained evolution of a three-dimensional Gaussian temperature distribution being subject to diffusion only (black diamonds) compared with the analytical solution (red solid line) at different times. The blue dashed line shows the relative error between both solutions. The temperature profile is shown along the x-direction.
2012.10535
cond_evol_2.jpg
0.44717
c26e4431c2ef457ea849750898fe6029
Evolution of Hα intensity as measured for the simulated reference clouds M0 (black solid line) and L0 (red solid line). The dashed-lines and dotted lines show the respective intensities at distances of 1kpc and 10kpc, respectively, in order to compare to <cit.>.
2012.10535
halpha_intensity_t.jpg
0.44227
aa2be47d3ef44359ab0701debac4bd75
The same as Fig. <ref>, but for the low-mass model clouds. The arrow lengths are scaled to v_ ini=248 (model L1b_vel).
2012.10535
hvc_dens_plane_xy_3.jpg
0.389286
8e6868be19774a509c8e501a5138eb1e
Evolution of density in the massive model clouds. The plots show the section plane (x,y) through cloud centre. The upper most plot shows the initial stage. The evolution is shown in snapshots after 20Myr (left column) and after 50Myr (right column). Model M3_hom_tc_sg is shown at 55Myr, because the finger-like structures, which are due to , are substantially developed. Model M1_dec is shown at 100Myr when the head-tail structure is more prominently developed. The black solid lines enclose cloud material (gravitationally bound and physically connected) and the light blue solid lines border gravitationally instable regions. Black arrows illustrate the motion of gas. The arrow lengths are scaled to v_ ini=333 (model M1c_vel). The grey square in the upper left corner is spanned by 10× 10 cells of finest numerical resolution (Δ x=1.3pc).
2012.10535
hvc_dens_plane_xy_4.jpg
0.485961
4575ebd3873a48a5a91eae0777bc5f39
The same as Fig. <ref>, but for respective clouds without thermal conduction.
2012.10535
hvc_mass_hist_3.jpg
0.437948
01167c2e07d24c658a881586b24ec441
Evolution of mass (upper plot) for reference clouds. The corresponding mass-transportation rates (middle plot) show both loss (Ṁ<0) and gain (Ṁ>0) of mass with the average values (dashed lines) indicating a mass loss. The relative mass loss (lower plot) is higher in low-mass clouds.
2012.10535
hvc_mdot_hist_2.jpg
0.460958
08a5019127f943f6bb980fa9de831877
Distributions of metallicity in the stripped material of massive clouds (upper panel) and low-mass clouds (lower panel) after 20Myr (left column) and after 50Myr (right column).
2012.10535
hvc_metl_hist_mass.jpg
0.497577
ef4f543439a442b1afae28e4a4f66122
Metallicity profiles weighted with the square of particle density along the direction of motion for the massive clouds (upper plot) and low-mass clouds (lower plot) after 50Myr.
2012.10535
hvc_metl_profile.jpg
0.441484
9e02bcd7bcb2482bac7cffae71c58c07
Column densities of Hi for selected massive clouds after 50Myr (M1_dec after 100Myr). The black isodensity contours represent log(N_Hi/)=18,19,… Left column: Actual N_Hi for the clouds. Right column: Diluted N_Hi as being obtained at a distance of 100kpc by a beam of opening angle 9arcmin. The black circle shows the area covered by this beam.
2012.10535
hvc_ncol_plane_xy_HI_5_countours.jpg
0.4946
bd6f5c40198243319ac82ee5f38b6bf8
Evolution of maximum particle density (upper plot) and mass (lower plot) of homogeneous and isothermal clouds with and without self-gravity.
2012.10535
hvc_nmax.jpg
0.448687
b3d85be5de1a466293badb8b7a9c9d62
Distributions of temperature in the stripped material of massive clouds (upper panel) and low-mass clouds (lower panel) with thermal conduction (red dashed lines) and without thermal conduction (black solid lines) after 20Myr (left column) and after 50Myr (right column). The standard deviation in every bin is too low to be visible in the plots.
2012.10535
hvc_temp_hist_mass_3.jpg
0.524737
9ece5a76543347208002b894f64cfdf8
Upper panel: Mean density of model clouds M0 (black solid line) and M1_me (red dashed line). Lower panel: Distributions of temperature within the clouds (left column) and in the stripped material (right column) after 20Myr (upper row) and after 50Myr (lower row). The standard deviation in every bin is too low to be visible in the plots.
2012.10535
hvc_temp_hist_mass_5.jpg
0.450071
affb6f2ad74142628429b2366dc8430f
Free-fall time versus disruption time in fastest massive clouds.
2012.10535
hvc_times.jpg
0.520236
2e710fd7275248cc99d2b95aa15695b7
From top to bottom: Initial radial profiles of density, temperature, pressure, and mass for all model clouds.
2012.10535
initialvalues2.jpg
0.463352
01d957ad98c8422fa233599bac4fbca4
Ratio of saturated to classical mass-transportation rate <cit.>.
2012.10535
mdotratio.jpg
0.5049
bd2ca510f8e7406d83e054381cd365d6
Radial cloud-to-Bonnor-Ebert mass ratios for all model clouds. The x-axis is normalized to the respective radius of each cloud.
2012.10535
prof_mbe.jpg
0.454673
43fdb5b93aae4c5381bbce3674a59444
Both disruption time τ_ dis and distance D_ dis for clouds developing a gravitationally instable region. The light-green stars indicate the onsets of collapse.
2012.10535
time_disruption_2.jpg
0.541475
a1fe436c4f0a459794ebf6ec8fd429ce
Deceleration of model clouds that experience drag.
2012.10535
vdot.jpg
0.439199
c2e79457e57f4bd1b256e1ae1177998e
PA concept in urban and rural areas with the vehicles served by the terrestrial BS and satellite, respectively.
2012.10537
Figure1_0814.jpg
0.554139
a953e097b6ac43588e043ef8e510fde2
E2E throughput of PA system in urban areas with spatially-correlated Rayleigh fading conditions. We set SNR = 21 dB, codeword length = 10^4 channel use, the minimum processing delay of the BS = 5 ms, antenna separation = 1.5 times the wavelength, and carrier frequency = 2.68 GHz.
2012.10537
Figure2_0910.jpg
0.41963
feefb69673cb46fa8c7e688c270544c4
Effect of spatial mismatch on the MRT and DFT BF in an NLoS multi-path propagation environment.
2012.10537
Figure3_0814.jpg
0.443605
7b511e21e78a4c41b8b12a7ef7cdbbbb
Received power at the vehicle for various BF schemes, prediction schemes (ideal prediction, without prediction, i.e., with spatial mismatch), and antenna sizes N, in a spatially-correlated Rayleigh fading environment.
2012.10537
Figure4_1217.jpg
0.453729
745905a9087840fea3f4aa657a107118
E2E throughput of the last wagon in a 10-wagon train served by the satellite, with SNR = 26 dB, codeword length = 10^4 channel use, and the minimum processing delay = 10 ms, antenna separation between the first and the last antenna in the same wagon is 10 times the wavelength, and carrier frequency is 2.68 GHz. The distance for adjacent PA and RA in closest wagons is 0.2 m.
2012.10537
Figure5_0814.jpg
0.392001
a7165da2504945f28b0ef9978959c423
Open symbols: Compilation of decoupling (freeze-out) times, τ_f, plotted as a function of the cubic root of the charged particle density, (dN_ch/dη)^1/3, observed in various collision systems and for a wide range of beam energies <cit.>. Filled symbols: extrapolated values of τ_f corresponding to each of the centrality classes used in our computation of η/s based on Au – Au collisions at √(s_ NN) =0.2 TeV and Pb – Pb collisions at √(s_ NN) =2.76 TeV, measured by the STAR and ALICE collaborations, respectively. Red/blue solid lines represent systematic uncertainties of the polynomial fits to the data.
2012.10542
Fig1.jpg
0.42156
da294b4c785a4feb82c46cc7b87b0a4f
Longitudinal width, σ_Δη, of the G_2^CI correlator vs. the estimated number of participants measured in Au – Au collisions at √(s_ NN) = 200 GeV <cit.> and in Pb – Pb collisions at √(s_ NN)=2.76 TeV <cit.>, reported by the STAR and ALICE collaborations, respectively. Error bars and error boxes represent statistical and systematic uncertainties, respectively.
2012.10542
Fig2.jpg
0.416456
7caba646109344828177d64fb7898519
Values of the shear viscosity per unit of entropy density, η/s, computed in this work, as a function of the cubic root of the charged particle density, (dN_ ch/dη)^1/3, measured in Pb – Pb collisions at √(s_ NN)=2.76 TeV <cit.> and in Au – Au collisions at √(s_ NN) = 200 GeV <cit.>. Error bars and error boxes represent statistical and systematic uncertainties, respectively.
2012.10542
Fig3.jpg
0.450413
c1d25d706dde4f53b5c8f503bb6ba307
Comparison of η/s obtained in this work with a collection of values published since 2000 <cit.>, including both theoretical values and experimental values obtained from direct comparisons of theory to data. Some more details are given in the text.
2012.10542
Fig4.jpg
0.418976
0c2c853f36634d2dac40b655ebbcb209
2012.10544
backdoor-code.jpg
0.623664
ea0d1765c4544a8694900c5c349fd303
2012.10544
backdoor-square.jpg
0.426311
4feb316cc9ef4a268778ddd9e41f1f21
In Panel (a), the tunneling (phase) time through the structure air(LR)^7air, whose transmission coefficient is shown in the lower panel of Figure <ref>. The tunneling times of the wave packet (WP) peaks through the air(LR)^7air structure, obtained from τ=∂ t_aSa/∂ω, at the resonant frequencies, are ≅ -0.28fs. In b), a Gaussian wave packet with the centroid at the resonance with frequency ω≃ 5.32 × 10^15rad/s, and the envelope as shown here is prepared, at t=0, at a distance z_0=L=40l_c from the SL.
2012.10546
Fig10e.jpg
0.448874
54f70d68461d439cbb701231daa4794b
A Gaussian WP with the centroid at the resonance with frequency ω≃ 5.32 × 10^15rad/s is prepared, at t=0, at a distance z_0=L=40l_c from the SL (see the blue curve). The WP moves towards D, passing through the SL (LR)^6, and reaches the point D (see the red curves), as predicted, at t=τ_D=2z_0/v_g+τ≃ 54.02fs. The WP is partially transmitted and partially reflected. Not all frequencies are transmitted because the transmission coefficients at the center of the packet are larger than in the tails. Thus, a dip is formed in the reflected (red) WP. The distance between the peak of the (blue) WP and the dip in the reflected (red) WP depends on τ.
2012.10546
Fig11e.jpg
0.45091
0d020de56cc749a9be513fe7efa13b2c
Incoming, reflected, and transmitted fields at the interface dielectric-conductor. b) The incidence θ_i and the effective refraction angle ψ for a silver slab with the dielectric constant shown here.
2012.10546
Fig1e.jpg
0.438625
b0b13a8ae05e4f188a92b1af3d203165
Transmittance as a function of the frequency through a single slab of silver with thickness d_c=30nm (left-hand side graph) and d_c=80nm (right-hand side graph), for incidence angles θ_i=π/6, π/4, π/3, and slightly less than π/2. For the slab thicknesses considered here, the EM field is highly attenuated below the plasma frequency ω_p= 5.75 10^15rad/s and oscillating for ω > ω_p. Near the incidence angle of π/2, the narrow and isolated resonances correspond to localized surface plasmons. Notice that as the incidence angle grows, the resonances move towards the isolated surface plasmon resonances.
2012.10546
Fig2e.jpg
0.482448
ac8a2e3c1ce144b88e0b4c2123219684
Transmittance through a single slab as a function of the incidence angle θ_i, for three different EM frequencies. For the graphs in the left-hand side column, the silver slab thickness is d_c=80nm, and in the right-hand side column, d_c=1000nm. Notice that the number of oscillations at high frequencies is of the order of d_c/λ.
2012.10546
Fig3e.jpg
0.465238
cd3e3bcf80f845c4b57f22b01b3654df
Transmittance as a function of the frequency through the metallic superlattices (air/silver/air)^n. In a) and b), the thickness of the silver and air slabs is equal to d_a=100nm and d_c=10nm. In a), the transmittance is shown for n=8 and incidence angles θ_i=π/6, π/4, π/3, and slightly less than π/2. Below the plasma frequency ω_p= 5.75 10^15rad/s, the transmittance is an oscillating function of ω. Near the incidence angle of π/2, the transmittance is highly resonant. In b), we repeat the transmittances for θ_i=π/6, π/4, but now for n=16, and we plot also the Kramer condition (black curves) and the dispersion relation of Equation (<ref>) derived in the theory of finite periodic systems (TFPS) (red lines). It is clear that this recurrence relation predicts the bands and the frequencies of all the resonant plasmons.
2012.10546
Fig4e.jpg
0.455987
31aca98afd244ea8b85789d41fa792d7
Transmittance as a function of the frequency through the metallic superlattices (air/silver/air)^n where the silver and dielectric widths are a) small and equal and b) different with the silver slabs' width being larger. The transmittances are shown for different incidence angles θ_i indicated on the graphs. Below the plasma frequency ω_p= 5.75 10^15rad/s, for small widths, the transmittance is independent of the incidence angle, while for larger silver width, the metallic superlattice is almost completely transparent. For frequencies above ω_p, we have wider bands for small layers' widths and thinner bands for larger silver widths.
2012.10546
Fig5e.jpg