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
|
---|---|---|---|---|
0.438618 |
9699bcd142a94a0db94ce076a9a40a59
|
Frequency distribution of the model of the TSS in a 10×10 lattice. We numerically simulate the first model of the TSS (Eqs. (<ref>), (<ref>)) in a smaller lattice than that considered in Fig. <ref>. Here, we arrange 10×10 sites and consider the open boundary condition. 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 20×20 lattices (cf. Fig. <ref>), 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=0.2.
|
2012.09479
|
figure22.jpg
|
0.425397 |
ac06ce343a344d30a84c397610045325
|
Amplitude distributions of the Lyapunov vectors. The absolute values of the first components of the Lyapunov vectors at each site are plotted in 10×10 square lattices. In panel (a), the Lyapunov vector associated with a positive Lyapunov exponent (the index n=50) is shown. One can confirm that the bulk sites exhibit large amplitudes. In panel (b), the Lyapunov vector of the index n=700 is shown. We can see its localization to the edge of the system. In panel (c), the Lyapunov vector associated with the smallest Lyapunov exponent (the index n=800) is shown. It is strongly localized to the upper-left corner. Thus, most of the sites have small amplitudes.
|
2012.09479
|
figure23.jpg
|
0.489377 |
f4eb1830a1f74f4ea43b3eb5038127f7
|
Dispersion relation of the Hamiltonian of topological lasing modes. We calculate the dispersion relation of the Hamiltonian of topological lasing modes utilized in our model (<ref>). We consider the open (periodic) boundary condition in the x (y) direction and arrange 20 sites in the x direction. We obtain doubly-degenerate gapless edge modes. Thus, the number of topological modes is four in this Hamiltonian, which is half of that in the state-dependent Hamiltonian. The parameter used are u=-1 and b=0.5.
|
2012.09479
|
figure24.jpg
|
0.418782 |
600414f5425241d5b5ddb06c6dfbe2cf
|
Band structure of lasing edge modes protected by conventional bulk topology. The band structure of the Hamiltonian of lasing edge modes (Eqs. (<ref>), (<ref>)) is shown. There exist gapless edge modes with larger imaginary parts of eigenvalues than those of the bulk modes. The parameters used are u=-1, u'=0.02, a=2, and b=0.5.
|
2012.09479
|
figure25.jpg
|
0.406161 |
de33434a97e24490983e5511c5392092
|
Absence of synchronization pattern in a topologically trivial parameter regime. We numerically simulate the first model of the TSS (Eqs. (<ref>), (<ref>)) in a topologically trivial parameter regime and confirm the absence of the TSS and its synchronization pattern. We damp the oscillators encircled by the green boxes. The parameters used are u=-3, b=0.3, α=0.5, β=1, ω_0=0.1, and Δω=0.2. In panel (a), the frequency of the first component of the oscillator at each site at time t=200 is depicted. The undamped oscillators exhibit inhomogeneous frequencies around -2.0. In panel (b), the indicator of the local order parameter of the first component of the oscillator at each site is shown. The indicator is small at every site, which indicates the absence of the synchronized oscillators.
|
2012.09479
|
figure26.jpg
|
0.412453 |
dba779f481974924b8ae574623773209
|
Lyapunov spectrum and the indices of the localization of the Lyapunov vectors. (a) Lyapunov exponents calculated from the first model of the TSS (cf. Eqs. (<ref>), (<ref>) and Fig. <ref>). They are arranged in descending order. The index of the Lyapunov exponent is rescaled for the maximum to be unity. The dashed horizontal line corresponds to zero Lyapunov exponents. Some of the Lyapunov exponents are positive, i.e., placed above the dashed line, which indicates the chaotic behavior of the oscillators. The parameters used are u=-1, b=0.5, α=0.5, β=1, ω_0=0.2, and Δω=0.2. (b) Proportions of the edge amplitudes P_ edge of the Lyapunov vectors in the TSS. We set the relative index of the Lyapunov vector to be the same as panel (a). The edge proportions rise steeply around the relative index 0.8. Therefore, the Lyapunov vectors of the small indices spread in the bulk, while those of the large indices are localized at the edge of the system. This increase accompanies the decrease of the Lyapunov exponents in panel (a), which indicates that the edge localized Lyapunov vectors are dissipative modes. (c) Inverse participation ratios (IPRs) of the Lyapunov vectors. We set the relative index of the Lyapunov vector to be the same as in panel (a). The IPRs show a small rise around the relative index 0.8, corresponding to the edge localization of the Lyapunov vectors. The IPRs also increase steeply at a larger relative index, which indicates that a few Lyapunov vectors are strongly localized at a few edge sites.
|
2012.09479
|
figure3.jpg
|
0.456981 |
3b58117d065d475cbf5629bee684a2d2
|
Dispersion relations and the dynamics demonstrating a spontaneous nonlinearity-induced boundary. (a) Edge dispersions of the Hamiltonian H̃ of topological insulator laser utilized in our first model under the existence of edge-localized on-site loss. We impose the open (periodic) boundary condition in the x (y) direction. The edge-localized on-site loss corresponds to the effect of nonlinear terms in the case that the edge modes of the original Hamiltonian H are amplified. We find the extra number of gapless modes. Those extra topological modes appear due to the effective boundary created by the nonlinearity-induced edge-localized loss. The parameters used are u=-1, b=0.8, α=0, β=1, ω_0=0, and Δω=0. (b) Density distribution of one of the extra topological modes corresponding to the red point in panel (a). One can see that this eigenstate exhibits the largest amplitude at the second site, which implies that it is localized at the effective boundary emerging between the first and second sites. (c) Dynamical formation of the extra topological boundary modes. We set the coefficient of the nonlinear term much smaller than that of the linear coupling. The snapshots at the time t=10 (t=40) are shown in the left (right) figure. When the edge oscillators are not fully amplified (corresponding to t=10), the edge-localized loss is not large enough to create an effective boundary. After the edge oscillators are fully amplified, the inside oscillators begin to largely oscillate, which implies the appearance of the extra topological boundary modes.
|
2012.09479
|
figure4.jpg
|
0.382391 |
6c23786998d448b2969e740ab9f9a1c3
|
TSS induced by conventional topological edge modes. (a),(b) Frequency of the first component of oscillators at each site. We numerically calculate the dynamics of the model of the TSS (<ref>) with different topological linear couplings (<ref>) from those in Fig. <ref>, where the edge modes are protected by conventional bulk topology. Panels (a) and (b) are the snapshots at the time t=1000 and t=2000 for each. Empty sites represent the oscillators oscillating around their natural frequencies. We confirm the coexistence of the frequency-synchronized edge oscillators and the desynchronized bulk oscillators. The parameters used are u=-1, u'=0.02, a=2, b=0.5, α=0.5, β=1, ω_0=1, and Δω=0.2. (c) Amplitudes of the oscillators obtained from the numerical simulation of the dynamics of the model considered in panels (a) and (b). We check that the amplitudes of most of the edge oscillators are larger than those of the bulk oscillators. The same parameters as panels (a) and (b) are used in the calculation.
|
2012.09479
|
figure5.jpg
|
0.461765 |
537d0d2a62ef4ec3b0df735be22f5ff3
|
Lyapunov analysis of the model of TSS utilizing lasing edge modes protected by conventional bulk topology. (a) Lyapunov exponents of the model with the linear couplings exhibiting edge modes protected by conventional bulk topology (<ref>). The index of each Lyapunov exponent is rescaled for the maximum to be unity. We obtain positive Lyapunov exponents (above the dashed line) as in Fig. <ref>, which indicates the chaos of the bulk oscillators. The parameters used are u=-1, u'=0.02, a=2, b=0.5, α=0.5, β=1, ω_0=0.2, and Δω=0.2. (b) Proportions of the edge amplitudes P_ edge (<ref>) of the Lyapunov vectors. We set the relative index of the Lyapunov vector to be the same as panel (a). The edge proportions rise steeply around the relative index 0.8, which indicates that the Lyapunov vectors of the small indices spread in the bulk, while those of the large indices are localized at the edge of the system. (c) Inverse participation ratios (IPRs) of the Lyapunov vectors. We set the relative index of the Lyapunov vector to be the same as in panel (a). The IPRs increase around the relative index 0.8, corresponding to the edge localization of the Lyapunov vectors. Since all the IPRs are less than 0.08, the strongly localized modes are absent unlike Fig. <ref>.
|
2012.09479
|
figure6.jpg
|
0.38805 |
d7b2bb9cbd684403accaa17322683681
|
Frequency and amplitude distributions of the model utilizing Hermitian linear couplings. We numerically simulate the model utilizing Hermitian linear couplings (<ref>). The frequency ((a) and (c)) and the indicator of the local order parameter M(x,y) (<ref>) ((b) and (d)) at each site are shown. Panels (a) and (b) ((c) and (d)) present the frequencies and the indicators of the local order parameters of the first (second) component of oscillators at time t=100. The edge oscillators at the first components of the lattice points exhibit the homogeneous and constant frequency and the large indicators of the local order parameters, which implies their frequency-synchronization. In contrast, the second components of oscillators are desynchronized. The parameters used are u=-1, b=0.5, α=1, β=1, ω_0=1, and Δω=0.2.
|
2012.09479
|
figure7.jpg
|
0.447349 |
df46911fbd4246a8b887ba5a544114ef
|
Lyapunov analysis of the model utilizing Hermitian linear couplings. (a) Lyapunov exponents of the model with the Hermitian linear couplings (<ref>) plotted in descending order. The index of each Lyapunov exponent is rescaled for the maximum to be unity. Positive Lyapunov exponents (above the dashed line) exist as in Fig. <ref>, indicating the chaos of the bulk oscillators. The parameters used are u=-1, b=0.5, α=1, β=1, ω_0=0.2, and Δω=0.2. (b),(c) Proportions of the edge amplitudes P_ edge (<ref>) of the Lyapunov vectors. We plot those of the first and fourth components in panel (b) and those of the second and third components in panel (c). We set the relative index of the Lyapunov vector to be the same as panel (a). The edge proportions of the first and fourth components rise steeply around the relative index 0.95, which indicates that the Lyapunov vectors of the large indices are localized at the first and fourth components of the edge sites. On the other hand, the edge proportion of the second and third components decreases in that region around the relative index 0.95. Therefore, the second and third components of the edge oscillators exhibit chaotic behavior.
|
2012.09479
|
figure8.jpg
|
0.497169 |
7b1bfbb9f8dc440e8b76da9caa65819b
|
Amplitude distribution of the lasing edge modes. The squared amplitude of the lasing edge mode of the linear couplings (<ref>) is shown. The wavenumber is k=0 and the eigenvalue is E=ib. Panels (a), (b), (c), and (d) show the amplitude distributions of the first, second, third, and fourth components for each. The amplitudes of lasing edge modes are localized at the first and fourth components. The parameters used are u=-1 and b=0.5.
|
2012.09479
|
figure9.jpg
|
0.417701 |
7f9aba0208c14cd393e772d7e9efafd3
|
a: Power spectrum |Ĵ(ω,z) |^2 of the heat flux timeseries shown in figure <ref> and normalised by ^2. b: Spatial variations of the power spectrum |Ĵ(ω,z) |^2 in a with depth through the fluid layer, shown for several frequencies, the lower two being typical of porous convection and the larger two typical of fluid convection. c: Vertical decay rate r of the energy of low frequency harmonics of the heat flux, for = 10^8 with varied by an order of magnitude from 10^-6 to 10^-5 as indicated in the legend. The dotted line gives the power law ω^0.75 for reference.
|
2012.09480
|
fig13.jpg
|
0.457587 |
57c6bdb5ff97483ab4d8550b9504bdf9
|
Snapshots of the temperature field at different Rayleigh numbers in two stable (top panels) and two convective (bottom panels) cases for the porous medium. The Darcy number is kept at = 10^-5.5. The colour scale is cut at 0.8 to enhance the contrast in the fluid layer.
|
2012.09480
|
fig3.jpg
|
0.531284 |
fcfe1d55362c4f4cb8b3376b486c8c2c
|
a: Nusselt number Ν = 2 √()<J> characterising heat transport across the two-layer system, together with predictions from <ref> for the porous-stable case <ref> (red dashed) and porous-unstable <ref> (black dotted and dot-dashed, with 𝒞_p = 0.85 and 𝒞_f as marked). The horizontal line marks Ν = 2. b: Root-mean-squared vertical velocity amplitude in the porous and fluid layers. The two grey lines indicate a scaling of (the characteristic speed in the porous layer). c: Ratio of the diffusive (J_d,p) and advective (J_a,p) fluxes to the total depth-averaged flux in the porous layer. In all three panels, = 10^-5.5 and the vertical line marks the threshold of porous convection estimated using <ref>.
|
2012.09480
|
fig5.jpg
|
0.529517 |
1b9ab661376d4a5a93a68b31ae365732
|
The Nusselt number N, extracted from the measured flux <J> and temperature contrast across the boundary layers via <ref>, as a function of the rescaled Rayleigh number = 2 Θ^β in the fluid layer (panel a and subscripts f) and in the porous layer (panel b and subscripts p). The values of Θ, the temperature difference across the boundary layers, is extracted from numerical data. Numerical data are sorted by equal values of the damping factor . In each panel, fits from studies of the single-layer case <ref> or <ref> are included (dashed grey) together with asymptotic predictions for →∞ as discussed in <ref> (dotted red). Panel b also includes data from simulations of a single porous Rayleigh-Bénard cell with an open top boundary on which the fit <ref> is based (red stars). All but two simulations had _p = _f = 1; the values of _p for these two simulations are given in the legend.
|
2012.09480
|
fig8.jpg
|
0.399128 |
0df48726fd6e4d589788912d24b5b15f
|
a & b: Values of and (see <ref>) extracted from simulations for different porous velocity scale . c: Schematic representation of two limiting temperature profiles in the limit of large Rayleigh numbers based on panel a and b. The porous Rayleigh number is large so that = 0.85 whereas varies from 1 in the weakly confined limit (≲ 1) to 0.5 in the strongly confined limit (≪ 1).
|
2012.09480
|
fig9.jpg
|
0.440215 |
9da7367d379f4a669f6682d109c396fe
|
Hierarchy of referential game agents implemented in ReferentialGym and their relationship to the more general Module class.
|
2012.09486
|
ReferentialGym-Tiny-Simpler-Modular.jpg
|
0.588409 |
d59228dd95da4ff4a0e69860a5277755
|
Schematic representation of the non-reciprocal photonic lattice studied in Section <ref>.
|
2012.09488
|
HatanoNelson.jpg
|
0.517213 |
b352a9c53802467babd3fd50672f8251
|
Noise-to-signal ratio, assuming N_1^in = 0, as a function of the photonic lattice site, for a chain defined by (<ref>), with t_d = t_c, N = 30, ϕ = π/2, and ω_d=ω_0. Continuous lines are an exact calculation, carried out by computing the output noise and signal by numerically calculating the matrix Q(ω) with Eq. (<ref>), and then using Eqs. (<ref>, <ref>, <ref>, <ref>).
|
2012.09488
|
NtSj.jpg
|
0.529015 |
02bf060158e344fead71027f361da3c5
|
Different definitions of a topological index depending on whether we focus on the (coherent) output signal S^out_j or the (incoherent) output noise flux N_j. Left: The amplification of a coherent input signal (in this example, at lattice site j = 1) is determined by the matrix Q(ω), evaluated at the frequency of the coherent drive, ω = ω_d. Following the connection with the SVD described in the text, topological amplification occurs at non-zero values of the winding number ν(ω_d). Right: The output noise, on the contrary, receives contributions from the propagation of incoherently generated photons at any frequency ω. N^out_j (assuming no significant noisy component in the input signal, N^in_1 = 0) is dominated by those frequencies for which ν(ω) ≠ 0 and, thus, topological amplification occurs.
|
2012.09488
|
SVsN.jpg
|
0.429508 |
c89970442ded4b6da73a5d47c5f1b16c
|
Average total gain at resonance as a function of the number of sites of the photonic chain (<ref>) as a function of W (standard deviation the distribution of local mode frequencies), with values t_d = t_c, and γ_p = 0.1. Averages are taken over 5 10^3 instances of diagonal disorder.
|
2012.09488
|
gain_disorder.jpg
|
0.47635 |
af6ab94b53e9404c86d313a8459a13b6
|
Mapping between non-Hermitian coupling matrices (<ref>) and a chiral topological insulator Hamiltonian (<ref>). The mapping relies on expressing the SVD as the eigensystem of an effective extended Hamiltonian H(ω). The existence of zero-energy states in the topological insulator picture leads to the appearance of zero singular values of the coupling matrix and, thus, to the phenomenon of exponential directional amplification.
|
2012.09488
|
main_idea.jpg
|
0.507245 |
aba646f3f79a4439bd312b9b82e625db
|
(a) Schematic representation of a photonic lattice consisting of a periodic system of coupled localized photonic modes. (b) Two sites in a generic photonic lattice, together with the input/ouput fields and couplings described in the main text: G_jl is a coherent, photon tunneling term, whereas γ^(p)_jl and γ^(d)_jl are dissipative gain and loss couplings, respectively.
|
2012.09488
|
photonic_lattice.jpg
|
0.439808 |
18e1ae5364534ae49728482a5936cbf6
|
Representation of a 1D photonic chain that exemplifies the topological amplification process. A coherent input field drives the first lattice site j = 1 and this signal gets multiplied by the zero-singular vector u^N. The output signal at site j=N is then proportional to the spatially inverted vector Π u^N and it gets amplified by a factor 1/s_N(ω_d).
|
2012.09488
|
topological_amplification_chain.jpg
|
0.413253 |
a7f067997b304ecfa765a45b7e8bc8e7
|
Albedo
|
2012.09491
|
ab.jpg
|
0.448942 |
bad84d4cdb024a46972b3f2e28116cf0
|
BG
|
2012.09491
|
bg.jpg
|
0.527159 |
a3953029019a4981bd935d872daa0625
|
3D
|
2012.09491
|
cmap.jpg
|
0.490532 |
3576836c1356405e84341b0fbfd17304
|
Depth
|
2012.09491
|
depth.jpg
|
0.411128 |
1548f23c53434384ae0cfe89864b1eca
|
GT
|
2012.09491
|
dwarpori.jpg
|
0.467254 |
a5e5a44d76ea4e7e8b165ee7baf41c2f
|
Normal
|
2012.09491
|
normal.jpg
|
0.468542 |
5d886e90f1f54995af46bfd29b9e583b
|
CT film
|
2012.09491
|
ori.jpg
|
0.42052 |
c898d96af836445088c09af87bae0288
|
UV
|
2012.09491
|
uv.jpg
|
0.414052 |
2d1657f1cdae48c4bcde9ac27065e13f
|
An illustration of the optimal switching rule.
|
2012.09493
|
Opt_strategy_eps.jpg
|
0.358311 |
13d47104d8124bc0a384a4086147ed2b
|
2012.09493
|
Opt_switch_time.jpg
|
|
0.51289 |
233222e3087d4dfaba5bb8620963ef09
|
2012.09493
|
autocorrelation.jpg
|
|
0.514061 |
d6ec6fe4250746fcb43fe3617af4c72a
|
2012.09493
|
electricity_price_time_series.jpg
|
|
0.442152 |
d2c9d9b15d944a6b948a643791b814bb
|
2012.09493
|
finale_statica_comparata_parametro_c.jpg
|
|
0.479358 |
a4a95abee7c24521996e418696e590fe
|
2012.09493
|
finale_statica_comparata_parametro_elle_concavita_alta.jpg
|
|
0.431772 |
521e7989746c4561b30069e024426188
|
2012.09493
|
finale_statica_comparata_parametro_elle_concavita_bassa.jpg
|
|
0.484084 |
b5f2afee4e8144dc94c9bfa34ad408e6
|
Dependency of x^* w.r.t. the incentive, k, for different choices of the parameter λ.
|
2012.09493
|
finale_statica_comparata_parametro_k.jpg
|
0.457227 |
4c81235ad09e466ea840685de19a0b53
|
2012.09493
|
finale_statica_comparata_parametro_lambda.jpg
|
|
0.406765 |
cd9814e1e19a4e269cf158ac897cfe5f
|
Dependency of x^* w.r.t. the volatility σ for different choices of the parameter λ.
|
2012.09493
|
finale_statica_comparata_parametro_sigma.jpg
|
0.456856 |
70f540c256e44fd7849c3f1716ddcbd9
|
2012.09493
|
fuel_price_time_series.jpg
|
|
0.397366 |
494db3140e1c497e9c4483fcdfe874eb
|
2012.09493
|
opt_switch_thr.jpg
|
|
0.494184 |
0a8024af91314051b86aa46ef09773be
|
2012.09493
|
partial_autocorrelation.jpg
|
|
0.436231 |
7c9eabf5643d446c8ddd07afbf2ebbbe
|
Unit distance opportunity cost time series. The data are considered with monthly frequency.
|
2012.09493
|
plot_X_t.jpg
|
0.440423 |
469a540a04014351828a89e1a21bf2f4
|
Simulation results for Parity check codes of length n= 2, 4 and 8 over Rayleigh Fading Channels for hard-decision decoding, proposed Flip Decoding (FD) and soft-decision decoding.
|
2012.09497
|
Flip_Decoder_dh2_n=248.jpg
|
0.455573 |
ace47c99b36f4627b3b76bbf48b71529
|
Schematics of the experimental apparatus. The microwave cavity (orange) is immersed in the uniform magnetic field (blue shaded region) generated by the magnet (crossed boxes). A1 and A2 are the cryogenic and room-temperature amplifiers, respectively. The JPA amplifier has three ports: signal (s), idler (i), and pump (p). Superconducting cables (red) are used as transmission lines for rf signals from the 4 K stage to the 150 mK stage. Thermometers (red circled T) are in thermal contact with the resonant cavity and the signal port on the JPA. Attenuators are shown with their reduction factor in decibels. The horizontal lines (blue) identify the boundaries of the cryogenic stages of the apparatus, with the cavity enclosed within the 150 mK radiation shield. The magnet is immersed in liquid helium.
|
2012.09498
|
ApparatoNicolo_ModDirectionalCoupler.jpg
|
0.436886 |
dd1fa14ded4b41f18900bb2fd22c157a
|
View of the QUAX-aγ dilution refrigerator insert, instrumented with the resonant cavity (at the bottom) and amplification chain. In the background, the 8.1 T magnet with its countercoil is visible.
|
2012.09498
|
SetupCrioLNL-JPA.jpg
|
0.473201 |
7d5cda73197d4b4090fa4769237197a8
|
The 90% single-sided C.L. upper limit for the axion coupling constant g_a γγ. Each point corresponds to a test axion mass in the analysis window. The solid curve represents the expected limit in the case of no signal. The yellow region indicates the QCD axion model band. We assume ρ_a∼0.45 GeV/cm^3.
|
2012.09498
|
ULgagg90CL_ML.jpg
|
0.46347 |
12597c0465ba4c94a87198f2289db680
|
Aggregate plot of the limits on g_aγγ obtained from the main axion search experiments. The two limits obtained by the QUAX Collaboration are highlighted in purple. The gray area identifies the region where axions could be found, with the yellow band and the two solid red lines identifying the coupling predicted by the KSVZ and DFSZ models and its uncertainty.
|
2012.09498
|
exclusion_plot.jpg
|
0.451516 |
165c09c2e77c4b5a860e8644b38d16b4
|
Measured power spectrum. The red line represents the model function used to obtain residuals.
|
2012.09498
|
fft_fit.jpg
|
0.468416 |
a7f5ed536598496eb05bb7d3d9a3c64c
|
Distribution of the residuals normalized to the expected thermal noise.
|
2012.09498
|
hRes_fit.jpg
|
0.456645 |
c1b2258fd4124a8282a5f1cb8fbb6a6f
|
l-V and l-b distributions of the 18,413 molecular clouds identified in the second Galactic quadrant.
|
2012.09500
|
LVdia.jpg
|
0.413667 |
dd23cdb645c14ab69276ea87de7d831d
|
l-V diagram of the largest molecular cloud G125.1+02.6.
|
2012.09500
|
Q2LV.jpg
|
0.466143 |
723bfdcd7fd747a1b9cf1967b21999d6
|
Comparison with maser-parallax-based distances derived from the model of <cit.>. The error bars represent the standard deviation, and the 5% systematic error is included for Gaia distances.
|
2012.09500
|
compareReid2019disQ2.jpg
|
0.465468 |
752c405c8d404187815421a3958e19c9
|
Face on view of 76 molecular clouds with accurate distance measurements in the second Galactic quadrant. The color represents the radial velocity, and the error bar the standard deviation of distance includes the 5% systematic error. The origin of the coordinate is the Galactic center, and the marker sizes are scaled with the mass.
|
2012.09500
|
faceOndisQ2.jpg
|
0.469603 |
dc81db6cb7644814a769178a0c4a76cf
|
Distances to the largest molecular cloud G125.1+02.6. The edge of the molecular cloud is delineated with the red line. The integrated velocity range is [-22.63, 7.55] . The distance is in pc, and distances to many molecular clouds around G125.1+02.6 are also displayed.
|
2012.09500
|
largestCloud.jpg
|
0.375625 |
8ea7128d8e5a4cefaa3473a56a4b2219
|
Visualization for the clean and adversarial examples. The numbers above the pictures represent the L_∞-norm of the perturbations.
|
2012.09501
|
visualize.jpg
|
0.387157 |
193569ed2316425696bb7410763af143
|
Dependence of the EPR uncertainty as given by the HP approximation on various parameters. In all the plots shown, k=1, r_1=0.9 and r_2=0.8. Other fixed parameters have the values indicated in each plot. The orange dots in the panels of the left column correspond to the fixed values of N_1/N_2 and ν_2/μ_1 used in the other two panels. The orange squares in the panels of the right column correspond to the “ideal” choice μ_1=μ_2 and ν_1=ν_2.
|
2012.09507
|
supergeneralEPR.jpg
|
0.443853 |
9c12409fbdf344bcb14fbbf1254293ad
|
Laser spectral (left) and temporal intensity (right) measured for different hollow core fiber pressures P_HCF=1100 mbar, 1000 mbar, 700 mbar, 250 mbar (from top to bottom). Legends provide the RMS spectral width σ (left), and FWHM duration τ_fwhm (right) for each P_HCF. The thin curves represent |E|^2, and the thick curves the envelope. All curves are normalized.
|
2012.09510
|
Figure1.jpg
|
0.49628 |
9afa9cd7e66e4638aa37706d018c7dde
|
Measured plasma profile of the one-sided shock jet in nitrogen with a 22 bar backing pressure (blue,solid) and plasma profile of the supersonic jet with a 25 bar backing pressure and assuming L-shell ionization (red, dashed) at 150 from the nozzle.
|
2012.09510
|
Figure2.jpg
|
0.512699 |
44090a9a3a17405eb5df74f3d89c1adb
|
(a) Electron angular distributions for different plasma density and laser pulse FTL durations. Each image presents the average of 10 to 1000 shots, and the colorscale is then normalized to its maximum level. On the right, the following beam parameters are plotted as a function of plasma density, for different FTL pulse durations: charge per shot (b), mean energy of the accelerated electrons (c), and the quality factor as defined in the main text (d).The error bars represent the standard deviation from the mean value due to statistical fluctuations.
|
2012.09510
|
Figure3.jpg
|
0.398107 |
8151f9ca82f449f8b33176f8bca642fc
|
Parameter space of the kHz LPA. Measurements are represented by red circles with sizes proportional to the beam charge. Dashed and dotted lines represent conditions in <ref>, and the gray colormap corresponds to the normalized plasma density N_p = n_p / γ_⊥ n_c.
|
2012.09510
|
Figure4.jpg
|
0.452067 |
ac23eb25f76f46108f5932251cd043ea
|
Measured and modeled electron beams with 3.5 fs laser pulse and n_p=1.8 × 10^20 cm^-3 (blue in (a), top in (b,c,d)), and with 10 fs pulse and n_p=3.5 × 10^20 cm^-3 (red in (a), bottom in (b,c,d)). (a) Measured electron spectra (solid curves shaded with standard deviation) and modeled (dashed curves). (b) Measured (right) and simulated (left) angular electron distribution. (c) Simulated evolution of the peak laser field ε_x = eE_x/m_e cω_0 (red curve) and the electron energy spectrum normalized at each time-step (blue) (blues colors); the dashed red curve is the in-vacuum laser propagation, and the gray area represents the plasma density profile in arbitrary units. (d) On-axis laser field (red curve), its temporal envelope (red filled area), and electron density normalized to 15n_p (gray scale) at the laser focal position, just before injection.
|
2012.09510
|
Figure5.jpg
|
0.435798 |
4ccb127509ea46fa94c352cdbe2cc3fa
|
a) Electron spectra monitored during the 5h run. Each spectrum is averaged on 100 shots. b) Typical beam measured before the run. The mean charge of 2.6 pC/shot is an average on 200 shots . c) Simulated (solid line) and experimental (dashed line) electron spectra at the beginning of the experiment (blue, simulation with an adjusted 3.35 mJ energy) and at the end of the experiment (yellow, simulation with 3.0 mJ )
|
2012.09510
|
Figure6.jpg
|
0.419886 |
bbe8a2a101a2464c9f6a4d7fb91bb0a5
|
Electron density (gray scale) in z-x plane, and in phase space zâϵ_z, where ϵ_z is the energy neglecting the contribution of momentum along x and y, for L-shell and K-shell electrons (blue and green respectively), longitudinal electric field (E_z) (red dashed curve), and laser envelope (red shaded area) for the a) : E_l=3.35 mJ case and b) E_l=3.0 mJ case. c) Mean position of injection behind the laser pulse and maximum longitudinal electric field (normalized to the value for E_l=3.8 mJ) according to the laser driver energy
|
2012.09510
|
Figure7.jpg
|
0.469154 |
e354c1cb587343ff89a6630cfeaf83ec
|
Nyquist plots for P3HT at different voltages in dark and fitted with equivalent circuit as shown in inset and the table shows the fit values to equation (<ref>).
|
2012.09524
|
dark_impedance.jpg
|
0.534609 |
a62210f3defe43639d3179fb2afe1cf4
|
(a) Real part of impedance vs frequency (b) Imaginary part of impedance vs frequency in dark. Data fits to equations (<ref>) and (<ref>).
|
2012.09524
|
dark_real_imaginary.jpg
|
0.475535 |
7bd72c0d10b24e97af66e865e8fce51d
|
Nyquist plots for P3HT at different voltages for doping-1 and fitted with equivalent circuit as shown in inset and the table shows the fit values to equation (<ref>).
|
2012.09524
|
doping1_impedance.jpg
|
0.515106 |
9a6daa7044fb40a59fd28dc22d06b5e8
|
(a) Real part of impedance vs frequency (b) Imaginary part of impedance vs frequency for doping-1. Data fits to equations (<ref>) and (<ref>).
|
2012.09524
|
doping1_real_imaginary.jpg
|
0.478058 |
fe9b59f29e214a26bf2bc10bf00fcfc1
|
Nyquist plots for P3HT at different voltages for doping-2 and fitted with equivalent circuit as shown in inset and the table shows the fit values to equation (<ref>).
|
2012.09524
|
doping2_impedance.jpg
|
0.460287 |
968f136f20f849249445aa77ed5b34a7
|
Nyquist plots for P3HT at different voltages for doping-2 and fitted with equivalent circuit as shown in inset and the table shows the fit values to equations (<ref>) and (<ref>).
|
2012.09524
|
doping2_real_imaginary.jpg
|
0.440071 |
a39212c6ef0647fd9973917a584d1b54
|
Nyquist plots for P3HT at different voltages in light and fitted with equivalent circuit as shown in inset and the table shows the fit values to equation (<ref>).
|
2012.09524
|
light_impedance.jpg
|
0.491318 |
de26f711043b43628644c1ff9793f123
|
(a) Real part of impedance vs frequency (b) Imaginary part of impedance vs frequency in light. Data fits to equations (<ref>) and (<ref>).
|
2012.09524
|
light_real_imaginary.jpg
|
0.501295 |
364b2a61e96f4377a3d623fd34dc539b
|
The calculated relaxation time from the circuit elements in the ordered regions as a function of the variation in carrier density and bias voltage.
|
2012.09524
|
relaxations.jpg
|
0.427691 |
37d2b0eea9b44c0da114faa1d583126b
|
XRD pattern of regioregular poly (3-hexylthiophene), inset shows the structure and the data.
|
2012.09524
|
xrd_p3ht.jpg
|
0.395782 |
0cfc131313ac4992be8bbd55b4764d80
|
(color online) Uncertainty on the reconstructed pump-probe phase due to the noise in the signals M_1 and M_2, for different values of Δ. Both graphs were plotted for δ M_1 = δ M_2 = δ M and A_1 = 2. a) Case where A_1 = A_2 (M_1 and M_2 have the same signal to noise ratio). b) An example with A_2 = A_1/2 (the signal to noise ratio is lower for M_2 than for M_1).
|
2012.09528
|
Figure_1.jpg
|
0.419171 |
3886124c0fff4cab8a88c9e71f526fa7
|
(color online) Sketch of the two-photon transitions involved in laser-dressed ionization. XUV and IR photons are represented with purple and red arrows respectively. The photoelectron spectrum (right part) shows the main harmonics peaks, spaced by twice the energy of the driving laser. Weaker sidebands are located between those peaks, at even multiples of the driving laser energy.
|
2012.09528
|
Figure_2.jpg
|
0.416339 |
a2f0fea1b53043559f56247e72fff48f
|
(color online) Electron recombination time calculated within the strong field approximation <cit.>, as a function of the harmonic order. We considered an Argon atom submitted to a 800 nm wavelength. We disregarded any envelope effect of the driving pulse. The short and long trajectories are represented with continuous and dotted lines respectively. The simulation was performed for different values of the IR laser electric field. The two triangles on the vertical axis indicate roughly the range of the XUV emission time in the case of short trajectories.
|
2012.09528
|
Figure_3.jpg
|
0.48004 |
caba0675c61646dc9cd1dc29837c7fbe
|
(color online) Experimental setup. BS: beam splitter, PM: plane mirror, TM: toroidal mirror. The interferometer arms are approximately 2.5 m long. The Cold Target Recoil Ion Momentum Spectrometer (COLTRIMS) experiment in the end station is a possible application of the LIZARD; it is not part of the stabilization system itself.
|
2012.09528
|
Figure_4.jpg
|
0.418684 |
f32648df16064c69a5f0507aa0f9e89a
|
(color online) Comparison between open loop (blue) and closed loop (orange) operation. a) Long term measurement of the phase error Δϕ(t) = ϕ(t) - ϕ_0. The inset shows the evolution of the delay stage position when the setpoint optical paths difference is changed by 100 nm, in the case of a non-optimized (dashed line) and optimized (continuous line) PID. As expected, the kick yields an overall 50 nm displacement of the stage (because of the mirrors corner cube configuration). For an optimized PID, we consider that the new setpoint value is reached in approximately 0.5 s. b) Phase error histogram. The full width at half maximum for the closed loop error is indicated with a red line. Assuming a gaussian distribution, it should be approximately 2√(ln 2)≈ 1.67 times the RMS error (224 mrad). c) Power spectral density of the phase error δϕ(t). The low frequency noise, corresponding to long term drifts, is efficiently suppressed by the LIZARD.
|
2012.09528
|
Figure_5.jpg
|
0.412595 |
603d1d84f2d74d23be830ad261f70941
|
(color online) Comparison between a classical open loop RABBIT scan (a,b,c) and a RABBIT scan stabilized by LIZARD (d,e,f). (a,d) Modulation of the photoelectron spectrum with the delay τ, after background subtraction and normalization. In the stabilized scan (d), Î_14 and Î_22 (white dotted bands) were used as the two modulated signals M_1 and M_2 for the active stabilization. (b,e) Î_16 (blue) and Î_20 (green), as a function of the delay. The signals are integrated over the width of the bands with the corresponding colors shown in a) and d). The black curves are sinusoidal fits. (c,f) Squared modulus of the Fourier transforms of Î_16 (τ) and Î_20(τ). The average value of each of the two modulated signal was first subtracted in order to remove very low frequency components. The vertical dashed line shows the theoretical 2ω_L sidebands angular frequency. The raw Fourier transforms (dotted lines) were smoothed using zero padding (solid lines).
|
2012.09528
|
Figure_6.jpg
|
0.406473 |
657d3ec1c10a45ba845923832483f086
|
(color online) Intensity correlation between four different sidebands extracted from the spectrum of Fig. <ref>, and sideband SB_22. Data corresponding to the open loop scan and to the closed loop scan are in blue and orange respectively. Black squares indicate the average positions of the sets of data corresponding to the same set-point phase. All axis have equal unit lengths. The increasing tilt of the ellipses is due to the attochirp. a) SB_20 versus SB_22. b) SB_18 versus SB_22. c) SB_16 versus SB_22. d) SB_14 versus SB_22. Those two particular sidebands were used for the LIZARD stabilization, they are in phase quadrature and yield a non-tilted correlation ellipse.
|
2012.09528
|
Figure_7.jpg
|
0.414982 |
83d208001dfa4ec1851ce60dd96eb7d4
|
(color online) a) RABBIT scan in argon stabilized by LIZARD that lasted approximately 2 hours. Between each 20 nm optical paths change, spectra were accumulated during 65 s, which corresponds to 6.5 × 10^5 laser shots. The yellow curve shows the 2 ω amplitude. The very sinusoidal oscillations of two of the sidebands (SB_18 and SB_20) are plot as an example in blue and green. (b-c-d-e) Spectral resolution of some sidebands phases (Rainbow RABBIT technique <cit.>), in the case of the 2 hours long stabilized scan (orange) and a fast (30 s) calibration scan recorded just before (blue). f) RMS noise of the spectral phase measurement, with respect to the number of laser shots over which the spectra were averaged during each scan step.
|
2012.09528
|
Figure_8.jpg
|
0.449157 |
eb5d6c4c63ce43e3af44e7ec3227ff4d
|
Screenshots from the game to demonstrate Recalculation.
|
2012.09539
|
recalc.jpg
|
0.40265 |
9c081663f856456db4c10e23cf1c8114
|
A screenshot of the game with color-coded shield display.
|
2012.09539
|
screenshot_snake_v4.jpg
|
0.41999 |
7e245b3a62b64ddfb10f9f55dd51d180
|
Schematic of the ROV-PV system in an Arctic environment. The 2D in-plane water velocity is estimated from the bubble motion, which is recorded with the ROV camera. The perforated hose can for example be suspended into the water by means of a pulley system, as shown here, or directly through a hole in the ice.
|
2012.09545
|
Hopen_drawing2.jpg
|
0.476318 |
d80bfee467334175948d17ef3f9cf624
|
Horizontal velocity profiles divided into 10 equally spaced vertical bins below wave crests. Wave frequency: 1.4, 1.6 and 1.8 Hz from upper to lower row, respectively. Wave amplitude: approximately 7, 14 and 21 mm from left to right column, respectively. Blue dots: observed horizontal bubble velocity u_b. Green error bars: mean and 2 σ uncertainty of the observed u_b within each bin (at least 10 observations must be present within the bin for the error bar to be displayed). Red line: theoretical horizontal water velocity u_w given by Eq. (<ref>). Orange shaded region: expected u_b from estimated bubble size and Eq. (<ref>). The observed mean u_b lies within the expected region for most of the vertical bins.
|
2012.09545
|
Profile_dimensional_run1-9.jpg
|
0.421595 |
67ad774ad34b4f96a8a486e79fee9e0f
|
A map of the area where the field work was carried out. The red dot indicates the location of the ice floe. Source: <cit.>.
|
2012.09545
|
Svalbard_map.jpg
|
0.439318 |
79578cf9e7cc42a1bbb26224eee949bb
|
Time series of surface elevation η obtained with ultrasonic gauges for 1.4 Hz monochromatic waves. The amplitude was approximately 7, 15 and 23 mm for top, middle and bottom panel, respectively. The shaded area marks the 10 periods investigated, which were periodic, almost constant in amplitude and unaffected by any reflected waves from the beach.
|
2012.09545
|
Time_series_eta_1p4Hz.jpg
|
0.468191 |
390d59adb03747f7b1e3a37121cb8166
|
Schematic drawing of the experimental setup in the wave tank. A coordinate system is defined with the (x, y) axes aligned horizontally in the direction of wave propagation and vertically in upward direction, respectively, with y = 0 as the undisturbed free surface. The velocity under the waves is estimated from the motion of the bubbles, which is recorded with the ROV camera.
|
2012.09545
|
exp_setup_waves-crop.jpg
|
0.46398 |
c05af34cafbe42ea8e872b3e7acc178a
|
Schematic of the experimental setup to determine the terminal velocity of rising bubbles in stagnant fluids with (x, y) axis aligned horizontally and vertically, respectively. The bubble motion was recorded with a high speed camera and the bubble velocity and size were found from image processing techniques.
|
2012.09545
|
oppsett5.jpg
|
0.454327 |
7fd13f58194949fc9f8c48abccdec535
|
(a) The work W(τ) calculated from Eq. (<ref>) and (b) the change of the von Neumann entropy of the main system Δ S_A^vN(τ) calculated from Eq. (<ref>). Both are plotted as functions of time τ/T for fixed η=1 and βħω_0 = 1. The black dots represent the results from the quasi-static distribution ρ̂_A^qeq(τ) that satisfy W^qeq(τ)=Δ F_A (τ). The colored curves represent different time durations: T=0.1/ω_0 (blue curve), T=1/ω_0 (green curve), and T=10/ω_0 (red curve).
|
2012.09546
|
Fig1.jpg
|
0.440706 |
bfad2706870d4d2ab2aba674e80afef8
|
(a) Change of the system entropy and (b) the total entropy production in the Boltzmann case (red curves) and the von Neumann case (blue curves) as functions of the system–bath coupling strength η, for the inverse temperatures: (i) βħω_0 = 0.5, (ii) βħω_0 = 1, and (iii) βħω_0 = 3.
|
2012.09546
|
Fig2.jpg
|
0.410384 |
d3f921a59acb4db2a72948456f737375
|
Change of the heat Δ Q (green curve), the decrease of the bath energy -Δ⟨ℋ̂_B⟩ (red curve), the change of the interaction energy Δ⟨ℋ̂_I⟩ (blue curve), and the change of the system part of the system–bath interaction energy δ U_A' ≡Δ U_A - Δ⟨ℋ̂_A⟩ (black dashed curve), plotted as functions of η for (a) βħω_0 = 0.5, (b) βħω_0 = 1, and (c) βħω_0 = 3.
|
2012.09546
|
Fig3.jpg
|
0.437501 |
76c5327a8f0e494a94691c943b439d06
|
(a) Total entropy production (blue curve) and the change of the system entropy (red curve). (b) Change of the system energy (green curve), interaction energy (blue curve), and bath energy (red curve). These were calculated from the factorized initial condition in the von Neumann case as functions of the system–bath coupling for (i) βħω_0 = 0.5, (ii) βħω_0 = 1.0, and (iii) βħω_0 = 3.0, under the same physical conditions as in Figs. <ref> and <ref>.
|
2012.09546
|
Fig4.jpg
|
0.487492 |
ed621b5e12e9408cb7eaa142c950ad27
|
Noise Encoder
|
2012.09547
|
new_archs.jpg
|
0.516136 |
aff0266dc4d943fd8584fb8f53cf652b
|
A practical `decision tree' to measure tension, illustrating when each tension metric should be used.
|
2012.09554
|
Tension_tree_v2.jpg
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.