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
|
---|---|---|---|---|
0.392433 | 671a8b33744c4af082bc27b798057e42 | Gas density shown on large scales for simulation B inside a slice parallel to the stellar orbital plane. The black hole is represented by the white dot while the white segment indicates the scale used. The black arrows indicate the direction of the velocity field while the white contour corresponds to the surface beyond which the gas becomes unbound from the black hole with ε = v^2/2 - G /R >0. | 2012.12271 | density_large_plane_non_relativistic.jpg |
0.371763 | 03aa469e42d74b67a96bf3de94504e69 | Gas density shown on large scales for simulation A inside slices parallel (left panel) and orthogonal (right panel) to the stellar orbital plane that contain the black hole. The value of the density increases from black to white, as shown on the colour bar. The white segment indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. The black arrows indicate the direction of the velocity field while the white contour corresponds to the surface beyond which the gas becomes unbound from the black hole with a positive orbital energy ε = v^2/2 - G /R >0. The dotted orange circle located at R=150 indicates the spherical surface used to evaluate the outflow rate in Fig. <ref>. The dash-dotted yellow line represents the trajectory of a gas particle injected on a parabolic orbit with ϵ =0 at the lowermost edge of the cone of unbound matter launched from the intersection point. The solid green line shows the trajectory of an initially bound particle that gets ejected due to radiation pressure while it would follow the blue dashed line if it was moving on an entirely ballistic orbit. | 2012.12271 | density_large_relativistic.jpg |
0.469308 | dce33b4960764604906f6b5e1e25cf92 | Snapshots showing the gas density for simulation B in a slice parallel to the stellar orbital plane that contains the black hole at different times of t/ = 0.1, 0.2, 0.3, 0.4, 0.5 and 0.6. The value of the density increases from black to white, as shown on the colour bar. The white segment on the first snapshot indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. | 2012.12271 | density_plane_non_relativistic.jpg |
0.501874 | d07ba14802a840279f43dcee24ee2391 | Snapshots showing the gas density for simulation A in a slice parallel to the stellar orbital plane that contains the black hole at different times of t/ = 5 × 10^-3, 0.03, 0.1, 0.2, 0.4 and 0.6. The value of the density increases from black to white, as shown on the colour bar. The white segment on the first snapshot indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. On the first snapshot, the white dotted line displays the precessed trajectory of the returning stream. The red solid arrow represents the velocity of a given outflowing particle while the blue dashed ones show its net and expansion components. | 2012.12271 | density_plane_relativistic.jpg |
0.490883 | da14c4ef97ee459ba31e55916beba924 | Gas density in a slice perpendicular to the stellar orbital plane that contains the black hole when the discs have formed at t/ = 0.4 for simulation A (upper panel) and t/ = 0.5 for simulation B (lower panel). The value of the density increases from black to white, as shown on the colour bar. The black hole is depicted by the white dot. The white segments indicate the scale used and the orientation is given by the white arrows. | 2012.12271 | density_vertical_comparison.jpg |
0.500135 | 53a1ff9929cc4efa8432c1cb424714cb | Eccentricity distribution for simulation A at t/ = 0.4 (black solid line) and simulation B at t/ = 0.5 (red dashed line) considering the gas within a radius R ≤ 10 and a vertical distance |z|≤ 2 from the mid-plane where the disc is located. The normalization is chosen such that the areas under the curves are equal to one. | 2012.12271 | eccdist.jpg |
0.480888 | b8e38d2c557f4843ad6a039828871e4f | Vertical profile of different rates of energy variation evaluated as the disc forms at t = 0.2 by considering the gas contained between the radius R_ in = 8 and R_ out = 12 within an azimuthal angle Δφ = π/8 from the - direction. The black solid line denotes the rate of inward kinetic energy transport due to the infalling matter computed as Ė_ kin = Ṁ_ in v^2_ z/2, where the vertical component v_ z of the gas velocity and the vertical net inflow rate Ṁ_ in are obtained from the simulation. The red dashed line has its shape given by the thermal energy increase due to shocks in each vertical layers, not taking into account cooling by free-free emission. These contributions are added up for vertical distances in the range 0≤ z ≤ 3 to obtain the total heating rate, which is used to fix the peak of this curve. The dot-dashed blue line corresponds to the rate of outward diffusion obtained from Ė_ dif = (F_ r·) Δ A by multiplying the vertical component of the radiation flux by the surface Δ A = (R^2_ out-R^2_ in) Δφ. | 2012.12271 | edotvsz.jpg |
0.467425 | 882df6229b8a488a9706813240ea1b19 | Evolution of the unbound fraction for simulations A (black solid line) and B (red dashed line). It is calculated from ratio of the mass of unbound matter with a positive orbital energy ϵ = v^2/2 - G /R >0 to that of all the evolving gas. The horizontal grey dotted line denotes the unbound fraction f_ unb = 0.33 of the gas launched from the self-crossing shock in simulation A. For simulation B, all the gas is injected on bound orbits. | 2012.12271 | funbvst.jpg |
0.40781 | 89a57b28982a474fa7d5465d98565466 | Isotropic equivalent luminosity in a slice parallel to the stellar orbital plane that contain the black hole for simulation B at t/ = 0.5. It is computed from L = 4 π R^2 (F_ r·e_ r) using the projection of the radiation flux directly obtained from the code along the radial direction, which is divided by the Eddington luminosity L_ Edd = 4 π G c / κ_ s≈ 10^44. This ratio is also that of the accelerations due to radiation pressure and gravity along this direction. It increases from blue to red, as shown on the colour bar. The white segment indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. The black arrows indicate the direction of the radiation flux while the blue solid line denotes the location of the scattering photosphere at the current time. | 2012.12271 | luminosity_plane_non_relativistic.jpg |
0.4049 | 6a45d9a277bf4b2d9f97695debce4faf | Isotropic equivalent luminosity in slices parallel (left panel) and perpendicular (right panel) to the stellar orbital plane that contain the black hole for simulation A at t/ = 0.4. It is computed from L = 4 π R^2 (F_ r·e_ r) using the projection of the radiation flux directly obtained from the code along the radial direction, which is divided by the Eddington luminosity L_ Edd = 4 π G c / κ_ s≈ 4 × 10^44. This ratio is also that of the accelerations due to radiation pressure and gravity along this direction. It increases from blue to red, as shown on the colour bar. The white segment indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. The black arrows indicate the direction of the radiation flux while the blue solid lines denote the location of the scattering photosphere at the current time. The orange dashed curve denotes the estimated thermalization surface, where the properties of the emerging spectrum are set. | 2012.12271 | luminosity_relativistic.jpg |
0.42763 | a0e0ea44593c4a3181e86981cbeac6df | Evolution of the net inflow rate for simulation A through a sphere of radius R = 20 located outside the outer edge of the forming disc (black solid line) and the accretion rate onto the black hole (red dashed line) obtained from the particles crossing the accretion radius at R_ acc = 6. | 2012.12271 | mdotinvst.jpg |
0.539589 | 6213019751c344ac8b3eb25208f16bd5 | Evolution of the total outflow rate (black solid line) for simulation A obtained from the gas crossing a spherical surface at R=150 while moving outward. The red dashed line indicates the unbound component, which is obtained by only considering the debris with positive orbital energy ϵ = v^2/2 - G /R >0. The contribution to this outflowing unbound gas from debris injected from the self-crossing shock on bound trajectories with ϵ < 0 at injection is displayed with the blue dash-dotted line. The grey dotted line represents the rate of mass injection given by the fallback rate according to equation (<ref>). | 2012.12271 | mdotoutvst.jpg |
0.46827 | f97dc435ee204a8186c2e33233ebbb17 | Vertical profile of the integrated optical depth (black solid line) obtained at t = 0.2 from τ = ∫_z^z_ maxρκ_ s z', setting the upper bound to z_ max = 10. It is calculated by considering the gas contained between the radius R_ in = 8 and R_ out = 12 within an azimuthal angle Δφ = π/8 from the - direction. The grey dotted vertical line denotes the height z_ dis = 1.25 corresponding to the surface of the accumulated gas where heating by shocks is the largest, as can be seen from Fig. <ref> that focuses on the same region. The red dashed line corresponds to the ratio c/v of speed of light to gas velocity that we evaluate from the simulation. | 2012.12271 | tauvsz.jpg |
0.421274 | 1093fac7e7af4d7788414f8b280f6ae3 | Snapshots showing the radiation temperature T_ r= (e_ r/a)^1/4 for simulation B in a slice parallel to the stellar orbital plane that contains the black hole at different times of t/ = 0.1, 0.2, 0.3, 0.4, 0.5 and 0.6. The value of the temperature increases from purple to yellow, as shown on the colour bar. The white segment on the first snapshot indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. | 2012.12271 | temperature_plane_non_relativistic.jpg |
0.491591 | 24ba2cd227ea43fca7670376282bc43b | Snapshots showing the radiation temperature T_ r= (e_ r/a)^1/4 for simulation A in a slice parallel to the stellar orbital plane that contains the black hole at different times of t/ = 5 × 10^-3, 0.03, 0.1, 0.2, 0.4 and 0.6. The value of the temperature increases from purple to yellow, as shown on the colour bar. The white segment on the first snapshot indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. | 2012.12271 | temperature_plane_relativistic.jpg |
0.413912 | 4764a5187f9844eb9adb78b9b54b0e63 | Gas density in a slice parallel to the stellar orbital plane that contains the black hole at t/ = 0.4 for simulation A. The value of the density increases from black to white, as shown on the colour bar. The white segment indicates the scale used and the orientation is given by the white arrows. The black hole is depicted by the white dot and the grey circle represents the intersection point, from which matter is injected into the computational domain. The black arrows of equal length display the direction of the velocity field. The two curves represent the trajectories of two particles that end up getting accreted after joining the disc directly from the intersection point (solid blue line) and following an interaction near the mid-plane at larger radii (dashed orange line). | 2012.12271 | velocity_plane_relativistic.jpg |
0.431957 | 01932641822c4025b96c795c08e4bc70 | One-loop corrections to the ALP–boson couplings c_WW and c_BB involving Higgs exchange. The Higgs doublet is represented by a long-dashed line. The sum of these diagrams does not give rise to UV divergences. | 2012.12272 | Higgsloops.jpg |
0.483949 | fac7ad67a8134f92b40da55e62a246ba | One-loop diagrams accounting for operator mixing through Yukawa interactions and gauge interactions. | 2012.12272 | Runningloops.jpg |
0.477646 | d372e2d8fc484356b4147022cea3bd21 | Dependence of the phase-space functions g_00(r) (blue) and g_+-(r) (red) on the ALP mass (with r=m_π^2/m_a^2). | 2012.12272 | a3pi.jpg |
0.478434 | 2078de1d54814f7dbff53c9ff6599c1c | Contributions to the a→ gg decay amplitude involving the ALP–gluon coupling (left) and the ALP couplings to quarks (right). The ALP is drawn as a dotted line. The black circles indicate vertices deriving from the dimension-5 operators in the effective Lagrangian (<ref>). | 2012.12272 | atogluons.jpg |
0.478434 | ce7e39173fdf4cfe89e316b3d602291e | Examples of one-loop matching contributions to the ALP–boson couplings. These diagrams do not give rise to matching contributions when the form of the effective Lagrangian in (<ref>) is employed. | 2012.12272 | bosonmatching.jpg |
0.540292 | b3b8ae85fd874b9d8e5508aedb5c2770 | Examples of one-loop and two-loop diagrams contributing at the same order in perturbation theory if c_VV and c_F have similar magnitude. | 2012.12272 | cVVandcF.jpg |
0.472819 | 4651bf6a9efc42b6a081d00fc28e2f54 | Axion–electron coupling c_ee(m_t) in the DFSZ model for different values of tanβ=v_u/v_d and axion masses: m_a=1 keV (solid), 1 eV (dashed), 1 meV (dashed-dotted) and 1 μeV (dotted). The red curve depicts the coupling c_ee(Λ) at the high scale Λ=4π f. | 2012.12272 | cee.jpg |
0.49416 | 50150437a6e84a5a93feffa263d8d47a | Leading-order contributions to the a→γγ decay amplitude in the chiral expansion. The π^0γγ coupling is obtained from the Wess–Zumino–Witten term not shown explicitly in (<ref>). | 2012.12272 | cgammagamma.jpg |
0.49416 | 2c2fae259e2142ca933df3dc2f6df7e5 | One-loop matching contributions to the ALP–fermion couplings. In the second diagram (V_1 V_2)=(WW), (ZZ), (Zγ) or (γ Z). In the last two diagrams V=W,Z, but in the sum of all contributions only the W-boson graphs with internal top quarks (plus the corresponding graphs with Goldstone bosons) give rise to non-zero contributions. | 2012.12272 | matchingfermions.jpg |
0.49325 | a4a9fcf2f44d49b3804b1000cc0a629b | Distribution of the clusters as a function of redshift. The blue blank histogram represents the sample distribution without the redshift correction, while the red hatched histogram shows the cluster distribution obtained by considering the redshift correction Δ z /(1+z) ∼ 0.02. Objects with z>0.6, not considered in the cosmological analysis, are covered by the shaded area. | 2012.12273 | Nz.jpg |
0.423926 | a30f58343ffb41d09c2d8a210d8e3152 | Comparison of the constraints on S_8≡σ_8(Ω_m/0.3)^0.5 obtained, from top to bottom, from the joint analysis of cluster counts and weak-lensing in the AMICO KiDS-DR3 catalogue (black dot), from the results obtained by <cit.> (blue dots), <cit.> (red dots), <cit.> (green dot), <cit.> (brown dot), <cit.> (magenta dot), <cit.> (orange dot), <cit.> (cyan dot). Median, 16-th and 84-th percentiles are shown. | 2012.12273 | comparison_ext.jpg |
0.466594 | 8add116565bd4f18a77abfe6840d4a3d | Number counts from the AMICO KiDS-DR3 cluster catalogue as a function of the intrinsic richness λ^*, in the redshift bins z∈[0.10, 0.30], z∈[0.30, 0.45], z∈[0.45, 0.60], from left to right. The black dots represent the counts directly retrieved from the catalogue, where the error bars are given by the Poissonian noise. The blue solid lines represent the model computed by assuming the cosmological parameters obtained by <cit.> (Table 2, TT,TE,EE+lowE), while the red dashed lines show the results based on the WMAP cosmological parameters <cit.> (Table 3, WMAP-only Nine-year). Both in the Planck and WMAP cases, the scaling relation parameters and the intrinsic scatter have been fixed to the median values listed in Table <ref>, retrieved from the modelling. The grey bands represent the 68% confidence level derived from the multivariate posterior of all the free parameters considered in the cosmological analysis. | 2012.12273 | modelled_counts.jpg |
0.459763 | f47cfb0d1e44472692fe6ff625fd128d | The logarithm of the masses in units of (10^14 M_⊙ h^-1), logM̅_200, from the AMICO KiDS-DR3 cluster catalogue as a function of the intrinsic richness λ^*, in the redshift bins z∈[0.10, 0.30], z∈[0.30, 0.45], z∈[0.45, 0.60], from top to bottom. The black triangles represent the mean values of logM̅_200, given by the mean of the marginalised posterior obtained in the weak-lensing analysis, while the error bars are given by 1σ of the posterior distribution. The orange lines represent the median scaling relation obtained by modelling only logM̅_200, following the procedure described in <cit.>. The grey bands represent the 68% confidence level derived from the multivariate posterior of all the free parameters considered in the cosmological analysis described in Section <ref>. | 2012.12273 | modelled_masses.jpg |
0.391224 | 43ca607a91054ecd8eeaeaa83fc6587c | Constraints on Ω_ m, σ_8, α, β, γ, σ_ intr, derived in a flat ΛCDM universe by combining the redshift bins z∈ [0.1, 0.3], z∈[0.3, 0.45], z∈[0.45, 0.6] and assuming a minimum intrinsic richness λ^*_ ob, min=20 for cluster counts. The shown posteriors are marginalised also over Ω_ b, τ, n_ s, h, s, q, and δ_ b. The blue contours represent the results obtained from the joint analysis of cluster counts and weak-lensing masses, while the orange contours show the posteriors on α, β and γ, derived from the analysis including only the weak-lensing masses as in <cit.>. The confidence ellipses correspond to 68% and 95%, while the bands over the 1-D marginalised posteriors represent the 68% of confidence. | 2012.12273 | results_final.jpg |
0.477253 | 3aa2eb7c984f48fea2ad66e71e1e0d9c | (a) Schematic picture of Choi–Jamiołkowski isomorphism between MPO and MPS (b) (c) Diagramatic representation of transfer matrix and operator transfer matrix | 2012.12274 | TNdiag.jpg |
0.48796 | 75ef4443b60c4818953370b306580d79 | Main: plot of β(t) obtained by a numerical integration of Eq. (<ref>). The horizontal dashed line denotes β^ obtained by solving c(β)=0. Inset: plots of c(β) and χ(β) defined in Eqs. (<ref>) and (<ref>). For each β, c(β) and χ(β) was calculated by the operator transfer matrix formula defined in Eqs. (<ref>) and (<ref>). | 2012.12274 | betagammat.jpg |
0.425456 | 4d52ba6c1e20443d907834c589dd8828 | γ dependence of Rényi-2 divergece density s_2^ with β^ = β^ for several fixed γ t. Lines with circles correspond to the states in the middle of relaxation, and lines with squares correspond to the NESS. Dasshed lines connect points on γ=0.005 and the origin (γ=0,s_2=0) with a straight line. | 2012.12274 | fixt.jpg |
0.443653 | 60d407fe5e864728b4d37a25f037de5a | Time evolutions of Rényi-2 divergece density s_2^ defined in (<ref>) for several initial temperatures simulated by the TDVP algorithm with time-step Δ t=0.2. D denotes the bond-dimension of MPO. | 2012.12274 | gammats2.jpg |
0.436663 | 62f7027947a14a81a629cef30174eb1d | The k D2-branes stretched between an NS5'-brane on the right and the stack of D4-branes realize an 't Hooft operator with the charge corresponding to the anti-symmetric representation ∧^k V of U(N). An NS5'-brane on the left would give an operator corresponding to ∧^k V. The locations of the D6-branes in the x^6-direction do not affect the vev. | 2012.12275 | SQCD-branes-higher.jpg |
0.431938 | e2d4c2cd51c94ac0b6da2f48bced875f | Experimental investigations of quantum many-body scars. (A) Two-dimensional atom array subject to global Rydberg lasers with Rabi frequency Ω and detuning Δ. (B) A quasi-adiabatic ramp of Δ and Ω prepares an antiferromagnetic state |AF_1⟩ with sublattice A excited, and a detuning quench launches non-equilibrium dynamics. Atoms in |g⟩ are imaged in optical tweezers via fluorescence while atoms in |r⟩ (empty circles) are expelled and detected as atom loss. (C) The Rydberg population on sublattices A and B undergo periodic oscillations (Inset: geometry used here). | 2012.12276 | Figure1.jpg |
0.431799 | 19d52e32f341499db2fbe1dd5779fcf5 | Universal empirical description of scar lifetime. (A) Different geometries used in this study. The lifetime τ of the sublattice excitation difference depends strongly on the geometry. (B) As a function of coupling to blockade-violating states (∝Ω^2/V_0) and next-nearest-neighbor (NNN) interactions, the scar decay rate 1/τ displays a bilinear dependence (Inset: cross-section of the plane). Schematics depict regimes where the two different decay processes dominate. | 2012.12276 | Figure2.jpg |
0.466545 | ce44be17f6a242de8349572d9e516c49 | Emergent subharmonic locking and stabilization. (A) Pulse sequence showing state preparation and quench with Δ_q(t). (B) Scar dynamics on a chain during quench to fixed optimal detuning (bare) with lifetime τ_fixed, and time-dependent detuning (drive) with modulation frequency ω_m = 1.24 Ω and lifetime τ_drive. The drive increases the scar lifetime and changes its frequency to ω_m/2. (C) Scar lifetime and response frequency as a function of ω_m, showing a lifetime increase and subharmonic locking. (D) Dynamics of the entire Hilbert space measured with experimental snapshots (0.5 million total bit strings). The microstates of the constrained Hilbert space are ordered by n_A - n_B, or equivalently by Hamming distance from |AF_1⟩ <cit.>. Right subplots highlight |AF_2⟩ and a state with a domain wall |DW_1⟩. (E) Reduced density matrix of a single atom in a chain (numerics) shows that driving reduces the growth of entanglement entropy S_ent. | 2012.12276 | Figure3.jpg |
0.414084 | fffa2d528f9645cb8f3a5f273d953886 | Robustness of the subharmonic response. (A) Dynamics of sublattice population difference after quench, as a function of modulation frequency. (B) Fourier transform intensity |S(ω)|^2 of data in (a), showing a harmonic locking for ω_m < 0.8 Ω and a subharmonic locking for ω_m > 0.8 Ω. (C) Phase diagram of the subharmonic response |S(ω=ω_m/2)|^2 in chain data (left), chain numerics (middle) from perfectly initialized |AF_1⟩ without experimental imperfections, and honeycomb data (right). (D) Increase of subharmonic rigidity (see text) with increasing system size. | 2012.12276 | Figure4.jpg |
0.442984 | b03a3db2060042cd9fbc8325f37b4d11 | Signatures of a 4^th subharmonic response. (A) ⟨n|_⟩A - ⟨n|_⟩B in the presence of two different drives with modulation frequencies of ω_m = 1.83 Ω and 2.13 Ω, resulting in responses at a 4^th subharmonic of ω_fit = 0.458(4) Ω and 0.534(2) Ω, respectively. Data is on a 9-atom chain with V_0 / 2π = 32 MHz and drive parameters Δ_m = 1.75 Ω and Δ_0 = 0, which is a different drive parameter regime than those used in investigating a 2^nd subharmonic response. (B) ⟨n|_⟩A - ⟨n|_⟩B data for modulation frequencies from 1.51 Ω to 2.61 Ω with same parameters as A. (C) Fourier transform intensity |S(ω)|^2 of data in B, showing signatures of a 4^th subharmonic response (dotted white line) while seemingly not as robust as the 2^nd subharmonic response focused on in this work. | 2012.12276 | SuppFig_4thsub_v2.jpg |
0.425498 | 4d154e9ef3744c1cbc2abe865abcf22d | Stabilization of pure PXP model under drive (numerics). (Top) Sublattice excitation probability for undriven (left) and driven (right) PXP model. (Bottom) Entanglement entropy across midway bipartition for undriven (left) and driven (right) PXP model. Numerics are calculated for a 22-atom chain with periodic boundary conditions and the timescale is set by Ω/2π = 4.2 MHz. “Bare” is a conventional quench to Δ = 0 and “Drive” is a quench to Δ = Δ_0 + Δ_m cos(ω_m t), with drive parameters Δ_0 = 0.5 Ω, Δ_m = 1.0 Ω, and ω_m = 1.33 Ω. These plots show that the cosine drive allows to delay the onset of thermalization even for the “idealized” PXP model, which describes perfect nearest-neighbor blockade (V_0 = ∞) with no long-range interactions. | 2012.12276 | SuppFig_PXPdrive_v1.jpg |
0.466729 | 5d54471fb4564fc393fa78eaf9d9214e | Independent measurement of decay parameters α and β. (A) Measured decay rate as a function of coupling to blockade-violating states ∼Ω^2/(4 V_0), obtained by measuring at different Rabi frequencies Ω during the quench on one-dimensional 9-atom chains with a fixed V_0/(2π) = 5.9MHz. The linear fit (dashed line) is performed on the first 8 points, which correspond to Ω/V_0 < 0.5. (B) Measured decay rate as a function of next-to-nearest-neighbor interactions. We prepare 9-atom chains with a variable staggering angle between neighboring sites, keeping the nearest-neighbor interaction constant at V_0/(2π) = 17.1MHz (insets). All error bars are given by fit uncertainties. The values for α and β are consistent with the fit in the main text Fig. 2 within error bars. | 2012.12276 | SuppFig_alphabeta_v2.jpg |
0.432247 | 1926d701da944c3795288a294cf7fdf3 | Subharmonic stabilization as a function of modulation amplitude Δ_m. (A) Dynamics of sublattice population difference after quench as a function of modulation frequency, measured on a 9-atom chain with nearest-neighbor interaction strength V_0/2π = 120 MHz = 28 Ω/2π, detuning offset Δ_0 = 0.5 Ω, and modulation frequency ω_m = 1.28 Ω. The Δ_0 we choose here is commensurate with the optimal fixed-detuning quench Δ_q,opt on this lattice, so the Δ_m = 0 line corresponds to data for optimal undriven scars on this lattice. (B) Fourier transform intensity |S(ω)|^2 of data in A. Upon applying drive amplitude Δ_m ≈ 0.3 Ω, the scar lifetime dramatically increases and exhibits a rigid subharmonic response at ω = ω_m/2 = 0.64 Ω, independent of drive amplitude, before degrading at drive amplitude Δ_m ≈ 1.3 Ω. | 2012.12276 | SuppFig_deltam_v1.jpg |
0.463988 | 3a82ed74cc364d09b01174b99cc421b8 | Optimal fixed detuning during a fixed-detuning quench. Quenching from antiferromagnetic state |AF_1⟩ to various fixed detunings Δ_0 on a 162-atom honeycomb lattice with V_0 / 2π = 17.1 MHz and Ω/2π = 4.3 MHz. The optimal fixed detuning on the honeycomb lattice is calculated to be Δ_q,opt = 1/2 ∑_i,j > NN V_ij≈ 0.153 V_0. An optimum is experimentally observed here close to Δ_0 ≈ 0.13 V_0, consistent with expectations from Eq. (<ref>). | 2012.12276 | SuppFig_fixeddetuning_v2.jpg |
0.41773 | 7f5215e56dd54973a8b59c1501f9020f | Initial-state dependence on dynamics. Plotted here are fixed detuning quenches in a two-dimensional lattice (54-atom decorated honeycomb with V_0 / 2π = 9.1 MHz and Ω/2π = 4.2 MHz). With a |ggg...⟩ initial state (left) the sublattice populations quickly equilibrate. With an |AF_1⟩ initial state (right) the sublattice populations oscillate and equilibrate at a significantly slower rate, whose rate is dominated by imperfect blockade and NNN interactions as explored in the main text. | 2012.12276 | SuppFig_gggDec_v1.jpg |
0.454635 | fe86fe74c2254a2fa7d6924747b33f68 | Response to drive for different initial states. (A) (Left panels) Fixed detuning quench for |AF⟩ (top) and |ggg…⟩ (bottom) initial states, showing an initial-state dependence to the ensuing dynamics and equilibration time. (Right panels) Time-dependent quench for |AF⟩ (top) and |ggg…⟩ (bottom) initial states. The |AF⟩ state scars are prolonged and the individual sublattice response is synchronously locked to half the drive frequency, whereas the sublattice populations of the |ggg…⟩ state show small oscillations at the drive frequency (harmonic response). (B) Fourier transform intensity of the individual sublattices |S_A(ω)|^2 and |S_B(ω)|^2, averaged together. The |AF⟩ initial state (top) shows a strong subharmonic response and also a weak harmonic response (which disappears for |S_A-B(ω)|^2 as plotted in Figure 4B of the main text). The |ggg…⟩ initial state (bottom) shows a harmonic response but no detectable signatures of a subharmonic response. | 2012.12276 | SuppFig_gggdrive_v2.jpg |
0.476561 | 0199c5e572514dc39ff1f63a18b44776 | Subharmonic locking with square pulse modulation. (A) Pulse sequence with square pulse drive. (B) Scar dynamics during a quench to a fixed optimal detuning (bare), and a time-dependent detuning (drive) with modulation frequency ω_m = 1.24 Ω. The drive increases the scar lifetime and changes its frequency to ω_m/2. (C) Scar lifetime and response frequency as a function of ω_m, showing a lifetime increase and subharmonic locking. | 2012.12276 | SuppFig_squarepulse_v2.jpg |
0.387147 | b9efd1414f3b4eab96be6479d6d68256 | System-size dependence of the subharmonic response. Fourier transform intensity |S(ω)|^2 of ⟨n|_⟩A - ⟨n|_⟩B traces for a chain of varying system size. A prominent subharmonic feature emerges and becomes more robust as the number of atoms in the chain increases, signifying that the subharmonic response is a many-body effect. All data here is a chain with V_0 / 2π = 51 MHz, and with drive parameters Δ_m = Δ_0 = 0.55 Ω. | 2012.12276 | SuppFig_systemsize_v2.jpg |
0.41215 | 269309c924094433882a3648204112e1 | Entanglement entropy dynamics of the Rydberg chain for a half-chain bipartition. The parameters of the system are V_0/2π = 51 MHz, Ω/2π= 4.2 MHz. The time-dependent detuning amplitudes are Δ_0 = 0.55 Ω, Δ_m = 0.55 Ω. (A) Comparison of entanglement dynamics with harmonic detuning, optimal time-independent detuning Δ_q,opt, and zero detuning reveals more than two-fold decrease in rate of entanglement growth due to presence of the drive. Data is shown for a 50-atom chain. The detuning parameters are Δ_q,opt = 0.0173 V_0 and ω_m = 1.2 Ω. (B) Dependence of entanglement growth on the frequency of the drive for a 24-atom chain reveals an optimal modulation frequency that corresponds to the slowest rate of entanglement spreading. Inset: Time averaged entropy S_ent = 1/T∫_0^Tdt S_ent for T = 1.5 μs shows a clear minimum around ω_m/Ω≈ 1.225. | 2012.12276 | figSv2.jpg |
0.420619 | faa99b208d8d4bbfad405f4625fbacd3 | State-dependent subharmonic revivals in the pulsed drive model. (A) Many-body revivals under Floquet unitary U_F for τ=τ_c and varying ε = π - θ for a L=14 chain with periodic boundary conditions. Revivals were calculated by taking the average of |⟨AF|U_F(π + ε,τ)^2n|AF⟩|^2 for n=1,2,...,100. (B) Here, we depict the dependence of subharmonic weight on the rotation angle τ under H_PXP and the deviation ε from the perfect echo point θ = π, calculated for N=400 driving periods. We see oscillations persist to larger ε for τ near τ_c. | 2012.12276 | supp_final_figure1.jpg |
0.426899 | 13ba0bc7504e4583b8f7b2e0a0f94a2c | Numerical simulations of microstate population dynamics. We plot the microstate distribution for (top) cosine driving (Δ_0=Δ_m=0.55Ω) and (bottom) pulsed driving (ε=0.5) at driving frequency ω_m=1.15Ω for the 1D L=9 chain. The states are ordered by their Hamming distance from the |AF_1⟩ state. The right and left columns depict a decomposition into microstates of the two symmetric and anti-symmetric superpositions of the two Floquet eigenstates with largest overlap with |AF_1⟩, |AF_2⟩ states respectively. Dynamics (center column) appear to be largely explained by these two eigenstates, as can be seen from the agreement between microstate populations at stroboscopic times. We note that the microstates populated at stroboscopic times, for both the simulations involving cosine and pulsed driving, are in qualitative agreement with those observed for the experimental protocol, as shown in Fig. 3D of the main text. | 2012.12276 | supp_final_figure3.jpg |
0.464852 | 4210cefdb7b0452996cac935cf498c94 | Reconstructed pseudoscalar and scalar susceptibilities (in lattice units) in the I=1/2 channel from (<ref>) and (<ref>), respectively, with light- and strange-quark condensate data from <cit.>. In the inset panel we show χ_S^κ (T) separately in order to emphasize the peak behavior. For such lattice setup, the continuum extrapolation to the physical mass case m_s=27m_l gives T_c=154± 9 MeV. | 2012.12279 | Kkappa.jpg |
0.46244 | 5f293e067c574de8a89b7f11805e8393 | Probing correlations and coherence in the striated phase via quench dynamics. a. Unit cell of striated ordering (dashed box) with (0,0) and (1,1) sublattices outlined in red and blue, respectively. The fill shade on each site reflects the mean Rydberg excitation. b. The variational states for the (0,0) and (1,1) sublattices are illustrated on the Bloch sphere (see Methods). The black arrow illustrates the phase ϕ_q of Ω during the quench. c. Probability P^(d) of an excitation, conditioned on observing no nearest-neighbor excitations, and zero (red), three (light blue), or four (dark blue) diagonal next-nearest neighbor excitations. P^(0) is plotted for ϕ_q = π/2, showing resonant de-excitation of the (0,0) sublattice near the bare-atom resonance (leftmost vertical line). P^(3) and P^(4) are plotted for ϕ_q = -π/2, showing excitation peaks for the (1,1) sublattice at interaction shifts corresponding to 3 or 4 diagonal neighbors (two rightmost vertical lines). d, e. P^(0) and P^(4) vary with quench phase ϕ_q at their corresponding resonances (Δ_q/2π = 1.4 and 20.4 MHz, respectively), demonstrating coherence on both the (0,0) and (1,1) sublattices. Solid line fits are used to extract Bloch vector components. | 2012.12281 | FIG4_V4_new.jpg |
0.504084 | 543a698c21f14ff687c937472d4371b9 | Programmable two-dimensional arrays of strongly-interacting Rydberg atoms. a. Atoms are loaded into a 2D array of optical tweezer traps and rearranged into defect-free patterns by a second set of moving tweezers. Lasers at 420 nm and 1013 nm drive a coherent two-photon transition in each atom between ground state |g⟩ = |5S_1/2, F=2, m_F=-2⟩ and Rydberg state |r⟩ = |70S_1/2, m_j=-1/2, m_I = -3/2⟩. b. Fluorescence image of initial random loading of atoms, followed by rearrangement to a defect-free 15×15 (225 atoms) square array. After this initialization, the atoms evolve coherently under laser excitation with Rabi frequency Ω(t) and detuning Δ(t), and long-range interactions V_ij. Finally, the state of each atom is read out, with atoms excited to |r⟩ detected as loss and marked with red circles. Shown on the far right is an example measurement following quasi-adiabatic evolution into the checkerboard phase. c, d. Similar evolution on honeycomb and triangular lattices result in analogous ordered phases of Rydberg excitations with filling 1/2 and 1/3, respectively. | 2012.12281 | Figure1_v16.jpg |
0.385985 | 1cfc8c5345ae4792ab6273e696b7f259 | Benchmarking of quantum simulator using checkerboard ordering. a. A quasi-adiabatic detuning sweep Δ(t) at constant Rabi frequency Ω is used to prepare the checkerboard ground state with high fidelity. b. Two-site correlation function G^(2)(k,l), averaged over all pairs of atoms on a 12×12 array, showing near-perfect alternating correlations throughout the entire system. c. Exponential fits of rectified horizontal and vertical correlations are used to extract correlation lengths in the corresponding directions ξ_H and ξ_V. d. Histogram of many-body state occurrence frequency after 6767 repetitions of the experiment on a 12×12 array. The two most frequently occurring microstates correspond to the two perfect checkerboard orderings, and the next four most common ones are those with a single defect in one of the corners of the array. e. Probability of finding a perfect checkerboard ground state as a function of array size. The slightly higher probabilities in odd×odd systems is due to commensurate edges on opposing sides of the array. All data in this figure are conditioned on defect-free rearrangement of the array. | 2012.12281 | Figure2_v8_AI.jpg |
0.41074 | 17850e8278974b73bc55d35250c79be2 | Phase diagram of the two-dimensional square lattice. a. Example fluorescence image of atoms in the checkerboard phase and the corresponding Fourier transform averaged over many experimental repetitions ⟨ℱ(𝐤)⟩, highlighting the peak at (π,π) (circled). b. Image of atoms in the striated phase and the corresponding ⟨ℱ(𝐤)⟩ highlighting peaks at (0,π), (π,0) and (π,π) (circled). c. Image of atoms in the star phase with corresponding Fourier peaks at (π/2,π) and (π,0) (circled), as well as at symmetric partners (π,π/2) and (π,0). d. The experimental phase diagram is constructed by measuring order parameters for each of the three phases for different values of the tunable blockade range R_b/a and detuning Δ/Ω. Red markers indicate the numerically calculated phase boundaries (see Methods). e. The order parameters evaluated numerically using DMRG for a 9×9 array (see Methods). | 2012.12281 | Figure3_v10.jpg |
0.488582 | 7331689dd5f041599cce96a2f6b8324e | Large arrays of optical tweezers. The experimental platform produces optical tweezer arrays with up to ∼ 1000 tweezers and ∼ 50% loading probability per tweezer after 100ms of MOT loading time. a. Camera image of an array of 34×30 tweezers (1020 traps), including aberration correction. b. Sample image of random loading into this tweezer array, with 543 loaded atoms. Atoms are detected on an EMCCD camera with fluorescence imaging. | 2012.12281 | Figure_LargeArrays.jpg |
0.420196 | 648853af4957442b9efa81f9cc3ef39b | Rearrangement protocol. a. Sample sequence of individual rearrangement steps. There are two pre-sorting moves (1, 2). Move (3) is the single ejection move. Moves (4-6) consist of parallel vertical sorting within each column, including both upward and downwards move. The upper panel illustrates the frequency spectrum of the waveform in the vertical and horizontal AODs during these moves, with the underlying grid corresponding to the calibrated frequencies which map to SLM array rows and columns. b. Spectrograms representing the horizontal and vertical AOD waveforms over the duration of a single vertical frequency scan during a realistic rearrangement procedure for a 26×13 array. The heat-maps show frequency spectra of the AOD waveforms over small time intervals during the scan. | 2012.12281 | Figure_Rearrangement_v4.jpg |
0.427123 | c829b9fa04f04e39bdf173fb26c3677e | Correcting for aberrations in the SLM tweezer array. The aberration correction procedure utilizes the orthogonality of Zernike polynomials and the fact that correcting aberrations increases tweezer light shifts on the atoms. To independently measure and correct each aberration type, Zernike polynomials are added with variable amplitude to the SLM phase hologram, with values optimized to maximize tweezer light shifts. a. Two common aberration types: horizontal coma (upper) and primary spherical (lower), for which ∼ 50 milliwaves compensation on each reduces aberrations and results in higher-depth traps. b. Correcting for aberrations associated with the thirteen lowest order Zernike polynomials. The sum of all polynomials with their associated coefficients gives the total aberrated phase profile in the optical system, which is now corrected (total RMS aberration of ∼ 70 milliwaves). c. Trap depths across a 26×13 trap array before and after correction for aberrations. Aberration correction results in tighter focusing (higher trap light shift) and improved homogeneity. Trap depths are measured by probing the light shift of each trap on the |5S_1/2, F=2⟩→|5P_3/2, F'=2⟩ transition. d. Aberration correction also results in higher and more homogeneous trap frequencies across the array. Trap frequencies are measured by modulating tweezer depths at variable frequencies, resulting in parametric heating and atom loss when the modulation frequency is twice the radial trap frequency. The measurement after correction for aberrations shows a narrower spectrum and higher trap frequencies (averaged over the whole array). | 2012.12281 | Figure_TweezerAberrations_v4.jpg |
0.568611 | ccc31745ac414bb2936edf0f3eff4681 | Characterizing microwave-enhanced Rydberg detection fidelity. The effect of strong microwave (MW) pulses on Rydberg atoms is measured by preparing atoms in |g⟩, exciting to |r⟩ with a Rydberg π-pulse, and then applying the MW pulse before de-exciting residual Rydberg atoms with a final Rydberg π-pulse. (The entire sequence occurs while tweezers are briefly turned off.) a. Broad resonances are observed with varying microwave frequency, corresponding to transitions from |r⟩ = |70S⟩ to other Rydberg states. Note that the transition to |69P⟩ and |70P⟩ are in the range of 10-12 GHz, and over this entire range there is strong transfer out of |r⟩. Other resonances might be due to multi-photon effects. b. With fixed 6.9-GHz MW frequency and varying pulse time, there is a rapid transfer out of the Rydberg state on the timescale of several nanoseconds. Over short time-scales, there may be coherent oscillations which return population back to |r⟩, so a 100 ns pulse is used for enhancement of loss signal of |r⟩ in the experiment. | 2012.12281 | Figure_microwaves_v1.jpg |
0.417765 | e5a50246954e492ba59cc5e833c97427 | Optimization of data collapse. a. Distance D between rescaled correlation length ξ̃ vs. Δ̃ curves depends on both the location of the quantum critical point location Δ_c/Ω and on the correlation length critical exponent ν. The independently determined Δ_c/Ω (blue line, with uncertainty range in gray) and the experimentally extracted value of ν (dashed red line, with uncertainty range corresponding to the red shaded region) are marked on the plot. b. Our determination of ν (red) from data collapse around the independently determined Δ_c/Ω (blue) is consistent across arrays of different sizes. c-e. Data collapse is clearly better at the experimentally determined value (ν=0.62) as compared to the mean-field (ν=0.5) or the (1+1)D (ν=1) values. The horizontal extent of the data corresponds to the region of overlap of all rescaled data sets. | 2012.12281 | fig_collapse_distance_v1.jpg |
0.461827 | ad6cd2a878bc4b84aec37443909b16f2 | Observation of the (2+1)D Ising quantum phase transition on a 16×16 array. a. The transition into the checkerboard phase is explored using a linear detuning sweep Δ(t) at constant Ω. The resulting checkerboard ordering is measured at various endpoints. b. Example of growing correlations G^(2) with increasing Δ/Ω along a linear sweep with sweep rate s = 15 MHz/μs. c. Growth of correlation length ξ for s spanning an order of magnitude from 15 MHz/μs to 120 MHz/μs. ξ used here measures correlations between the coarse-grained local staggered magnetization (see Methods). d. For an optimized value of the critical exponent ν, all curves collapse onto a single universal curve when rescaled relative to the quantum critical point Δ_c. Inset: distance D between all pairs of rescaled curves as a function of ν (see Methods). The minimum at ν = 0.62(4) (red dashed line) yields the experimental value for the critical exponent (red and gray shaded regions indicate uncertainties). | 2012.12281 | fig_kzm_v7_16x16.jpg |
0.404149 | ce823c8a5532476697f4736782f0d1e0 | Coarse-grained local staggered magnetization. a. Examples of Rydberg populations n_i after a faster (top) and slower (bottom) linear sweep. b. Corresponding coarse-grained local staggered magnetizations m_i clearly show larger extents of antiferromagnetically ordered domains (dark blue or dark red) for the slower sweep (bottom) compared to for the faster sweep (top), as expected from the Kibble-Zurek mechanism. c. Isotropic correlation functions G^(2)_m for the corresponding coarse-grained local staggered magnetizations after a faster (top) or a slower (bottom). d. As a function of radial distance, correlations G^(2)_m decay exponentially with a length scale corresponding to the correlation length ξ. The two decay curves correspond to faster (orange) and slower (blue) sweeps. | 2012.12281 | fig_local_stagg.jpg |
0.49143 | b533777bd1de4964a8edec1d418e019e | Extracting the quantum critical point. a. The mean Rydberg excitation density ⟨ n ⟩ vs. detuning Δ/Ω on a 16×16 array. The data is fitted within a window (dashed lines) to a cubic polynomial (red curve) as a means of smoothening the data. b. The peak in the numerical derivative of the fitted data (red curve) corresponds to the critical point Δ_c/Ω = 1.12(4) (red shaded regions show uncertainty ranges, obtained from varying fit windows). In contrast, the point-by-point slope of the data (gray) is too noisy to be useful. c. Order parameter ℱ̃(π,π) for the checkerboard phase vs. Δ/Ω measured on a 16×16 array with the value of the critical point from b. superimposed (red line), showing the clear growth of the order parameter after the critical point. d. DMRG simulations of ⟨ n ⟩ vs. Δ/Ω on a 10×10 array. For comparison against the experimental fitting procedure, the data from numerics is also fitted to a cubic polynomial within the indicated window (dashed lines). e. The point-by-point slope of the numerical data (blue curve) has a peak at Δ_c/Ω = 1.18 (blue dashed line), in good agreement with the results (red dashed line) from both the numerical derivative of the cubic fit on the same data (red curve) and the result of the experiment. f. DMRG simulation of ℱ̃(π,π) vs. Δ/Ω, with the exact quantum critical point from numerics shown (red line). | 2012.12281 | fig_sus_v4.jpg |
0.466223 | b89303a993434d6f9fc2b1544659c33a | Generating homogeneous Rydberg beams a. Measured Gaussian-beam illumination on the SLM for shaping the 420-nm Rydberg beam. A Gaussian fit to this data is used as an input for the hologram optimization algorithm. b. Corrected and measured wavefront error through our optical system, showing a reduction of aberrations to λ/100. c. Computer-generated hologram for creating the 420-nm top-hat beam. d. Measured light intensity of the 420-nm top-hat beam (top), and the cross section along where atoms will be positioned (bottom). Vertical lines denote the 105-μm region where the beam should be flat. e. Using the measured top-hat intensity, a phase correction is calculated for adding to the initial hologram. f. Resulting top-hat beam after feedback shows significantly improved homogeneity. | 2012.12281 | figure_tophat_v1.jpg |
0.506264 | a0ce2ee1f0b24b4388edca4917929075 | (a)–(c) Laboratory-based XRD patterns, (d)–(f) temperature dependence of longitudinal resistivities ρ_xx, (g)–(i) magnetic-field dependence of transverse resistivities ρ_xy on the Mn_2CoAl films grown by MS (left), IBAS (middle), and MBE (right), respectively. | 2012.12282 | MCA_AXRD_Fig1.jpg |
0.416257 | 0d70244a533a44afa152d704ff903642 | HAADF-STEM and EDS mapping images of Mn, Co and Al for the MS-Mn_2CoAl film (a-d) and the IBAS-Mn_2CoAl film (e-h), respectively. (i) High resolution STEM image of the IBAS-Mn_2CoAl film, (j)(k) nano-beam diffraction images taken from the spot A and B shown in (i), respectively. | 2012.12282 | MCA_AXRD_Fig2.jpg |
0.392901 | a226241615344fb8bb9cd979b6d62c92 | (a) HAADF-STEM image, EDS mapping images of (e) Mn, (f) Co, and (g) Al for the MBE-Mn_2CoAl film. (b) High resolution STEM image of the MBE-Mn_2CoAl film, (c)(d) nano-beam diffraction images taken from the [100] and [110] directions of the MBE-Mn_2CoAl film, respectively. | 2012.12282 | MCA_AXRD_Fig3.jpg |
0.470248 | cd7640adb0074b3aa02299e4721bda81 | (a) Synchrotron XRD rocking curves of the 002 and 004 reflections for the Mn_2CoAl thin films grown by MS, IBAS and MBE. (b) Lattice constants of the Mn_2CoAl thin films determined by synchrotron XRD. The lattice constants of bulk XA and L2_1 structures, 0.584 nm and 0.577 nm, respectively, are shown as dotted lines. | 2012.12282 | MCA_AXRD_Fig4.jpg |
0.406944 | 0280cdd92aac4eb1ab0bcb428de3c37d | Atomic structure models of (a) Mn_2CoAl in the XA structure and (b) Co_2MnAl in the L2_1 structure. Simulated anomalous XRD patterns of Mn_2CoAl XA and Co_2MnAl L2_1 structures with disorder in (c) I_002 near the Mn K-edge (inset of (c) shows I_002/I_004) and (d) I_111 near the Co K-edge. | 2012.12282 | MCA_AXRD_Fig5.jpg |
0.426673 | d5efc2c8286f490c958b8f7ff4ca9295 | Experimental anomalous XRD patterns of Mn_2CoAl films grown by MBE, IBAS, and MS. (a) I_002 and (d) I_002/I_004 profiles near Mn K-edge. | 2012.12282 | MCA_AXRD_Fig6.jpg |
0.409413 | 03b98b8b94e9416e82455649b9e2ea51 | Calculated anomalous XRD profiles for the (a) Co-rich and (b) Mn-rich regions of a MBE-Mn_2CoAl thin film compared with experimentally obtained I_111/I_004 around the Mn K-edge. The calculations were performed based on XA and L2_1 structure models in both the Co-rich and Mn-rich regions as shown in Tables. <ref> and <ref>. | 2012.12282 | MCA_AXRD_Fig7.jpg |
0.491318 | eca7dc1290db446f881231e00e8a98c2 | Anomalous XRD profiles of I_111/I_004 and I_002/I_004 near the (a) Mn K-edge and (b) Co K-edge of the Mn_2CoAl film. Points with lines show the experimental results, while solid and dotted lines show calculated profiles based on the two phase models with the obtained atomic compositions and occupancies. L2_1B-type and the disordered L2_1-type structures were determined for the Mn-rich and Co-rich phases, respectively, by anomalous XRD of the Mn_2CoAl film grown by MBE. | 2012.12282 | MCA_AXRD_Fig8.jpg |
0.424528 | 19f6bbcb236343ee9a6fa4d8c449bfbb | Area selective composition analysis of the MBE Mn_2CoAl film in (a) the Co-rich region and (b) the Mn-rich region using STEM-EDS mapping. | 2012.12282 | MCA_FigS1.jpg |
0.470592 | e173919af2b44dfba1d5ef04ad18f526 | Calculated anomalous XRD results for the Co-rich region of Mn_2CoAl thin film compared with experimentally obtained (a) I_111/I_004 at Mn K-edge, (b) I_111/I_004 at Co K-edge, (c) I_002/I_004 at Mn K-edge, and (d) I_002/I_004 at Co K-edge. The calculations were performed based on XA and L2_1 structure models in the Co-rich region. | 2012.12282 | MCA_FigS2.jpg |
0.483377 | 4738f4661da74303acc069e31ad0c4f8 | Calculated anomalous XRD results for the Mn-rich region of Mn_2CoAl thin film compared with experimentally obtained (a) I_111/I_004 at Mn K-edge, (b) I_111/I_004 at Co K-edge, (c) I_002/I_004 at Mn K-edge, and (d)I_002/I_004 at Co K-edge. The calculations were performed based on XA and L2_1 structure models in the Mn-rich region. | 2012.12282 | MCA_FigS3.jpg |
0.49484 | ed0f2b9412f34f94a556ef4d8d59ee7d | Calculated anomalous XRD for mixed structure compared with experimentally obtained (a) I_111 at Mn K-edge, (b) I_111 at Co K-edge, (c) I_002 at Mn K-edge, and (d) I_002 at Co K-edge. The combinations were created using selected A-6 and B-5 models in the Co-rich region and another four D-6, D-7, E-1 and E-2 models in the Mn-rich region. | 2012.12282 | MCA_FigS4.jpg |
0.460673 | 4fc5a390571d47439ee2a110571c0440 | Calculated anomalous XRD normalized by 004 for mixed structure compared with experimentally obtained (a) I_111/I_004 at Mn K-edge, (b) I_111/I_004 at Co K-edge, (c) I_002/I_004 at Mn K-edge, and (d)I_002/I_004 at Co K-edge. The combinations were created using selected A-6 and B-5 models in the Co-rich region and another four D-6, D-7, E-1 and E-2 models in the Mn-rich region. | 2012.12282 | MCA_FigS5.jpg |
0.455375 | d4a1fbb5a90c4eab81cbf5baad74ec05 | Synchrotron XRD color-coded maps of the Mn_2CoAl thin film grown by MBE for the (a) 002, (b) 004, and (c) 111 reflections at the Co K-edge. The peak intensity over a sample region (ca. 3 mm × 5 mm) was measured to check the homogeneity of the thin film. | 2012.12282 | MCA_FigS6.jpg |
0.47114 | 8784a0556a4f4ad0a67a6c2bec5da4d9 | (a) Spin 1/2 degrees of freedom, Ŝ_i⃗, on the honeycomb lattice are subject to Heisenberg J Ŝ_i·Ŝ_i + δ and Kitaev 2KŜ_i^δŜ_i+δ^δ exchange interactions. Here δ=1(red), 2(green), and 3(blue) runs over the there bonds, and a_1 and a_2 correspond to the lattice vectors. (b) First (solid) and second (dashed line) Brillouin zones. (c) Average sign ⟨sign⟩ as a function of V for various angles φ. The figure includes the ground-state phase diagram with antiferromagnetic (AFM), Kitaev spin liquid (KSL), zig-zag (ZZ), ferromagnetic (FM), and stripy (SP) phases, as proposed in Ref. <cit.>. Here we have set the temperature to T=1 in units of A. | 2012.12283 | Fig1.jpg |
0.472449 | ee506be8f367480ba8bc52a1afd7aa40 | Comparison between the radial profile of the shape parameters using three different algorithm LSIM(Volume), LSIM(Semi-Major) and EVIM. | 2012.12284 | Algorithms.jpg |
0.42767 | 8a071e25b0264efea934147204119efb | Three dimensional orientation of the reduced inertia tensor from the DM and SH for one example of twisted halo. | 2012.12284 | DM_EigenVectors991.jpg |
0.40812 | ab9edd3979b14e61826be6b3034a7ade | Radial profile of ε for stars in the central (left), FoF group with no-truncation (middle) and FoF stars truncated above 150 kpc(right) panels. Counter-rotating farther out substructures may dominantly flip the sign of total angular momentum and thus seem to convert the disky structures. They must therefore be removed when we analyze the radial profile of the ε. | 2012.12284 | Epsilon_Distance_NEW1.jpg |
0.435491 | 8e030e9c7e0646e4a3f6dc9634b839e7 | 2D projection of the surface density of the mass in 3D thin shells (in units of 10^5 × M_⊙ kpc^-2 for a twisted SH. Evidently the halo is re-orienting at different radii. We have computed the projection along a fixed direction in space, z direction in the TNG coordinate. | 2012.12284 | NEW1_Star_9.jpg |
0.467832 | 2cfb71ab94dd475fa31f46bdba5c62a1 | 2D projection of the surface density (in units of 10^5 × M_⊙ kpc^-2 for a twisted-stretched SH. Starting from an oblate shape in 3D initially, the halo is stretching and becoming more spherical at larger radii. We have computed the projection along the x direction in the TNG coordinate. | 2012.12284 | NEW_Star_13.jpg |
0.385597 | d73ab42bfd4941c5be262e4072c8f702 | Three dimensional orientation of the reduced inertia tensor for a twisted-stretched halo. There is a quick rotation in the orientation of different eigenvectors radially. | 2012.12284 | SH_EigenVectors13.jpg |
0.46056 | 7f386e2e2d6c4a6a83f79c09d33f2667 | Three dimensional orientation of the reduced inertia tensor for a twisted halo. There is a smooth rotation in the orientation of different eigenvectors radially. | 2012.12284 | SH_EigenVectors9N.jpg |
0.428985 | dc361687896d470fb989cfa4a7b7a0df | The radial profile of the median and percentile of the shape parameters including the FoF group stars. It is evident that the radial profiles of the median and percentiles of shape parameters are fairly close between the Central(Cen) and substructures(FoF group). This implies that FoF group stars statistically behave the same in shaping the SH as the central stars do. | 2012.12284 | Shape_sqT_NEW_HALO.jpg |
0.425764 | dbb0d42a2a6748baad53d5e934becf13 | Median and 16(84) percentiles of shape parameters (s,q,T) using 3 different algorithms. We studied the shape using two different versions of LSIM and using EVIM. In LSIM(Volume), we compute the shape in local shells with an enclosed volume fixed while in LSIM(Semi-Major), we keep the semi-major axis fixed. | 2012.12284 | Shape_sqT_NEW_StellarHalo.jpg |
0.427963 | 0eaf61bc52c941dd8a3dc680ef3eb332 | Impact of changing the threshold of the stellar halo, ε, on the radial profile of the median and 16(84) percentiles of shape parameters. | 2012.12284 | Shape_sqT_Star_Epsilon.jpg |
0.443732 | 0130ec0d29ad4d268ad133f88b9e0cd5 | Logarithm of the projected number density map (in units of kpc^-2) of stellar halo for a sample of 4 MW like galaxies from our galaxy sample in TNG50. Every row presents one galaxy with an ID number. From the left to the right, we zoom-in more on the central part of the halo. We have chosen stars with |ε| ≤ 1. | 2012.12284 | Stellar_HaloNN.jpg |
0.466452 | bc205d1a6d7042ed868f4c16e23e7ace | The radial profile of the Axes/r as well as the angle of min-inter-max eigenvectors with few fixed vectors in space. (upper) an example of twisted halo. (Bottom-panel) an example of twisted-stretched halo. | 2012.12284 | Stellar_Twisted_Stretched1.jpg |
0.448329 | 9ebc5cd97ed74c9d9f839a7e30c53e33 | The radial profile of Axes/r, angles and shape parameters for the twisted-stretched halos. | 2012.12284 | Twisted-Stretched1N.jpg |
0.427606 | 24250bd69ee4406d802f865151b959cb | The radial profile of Axes/r, angles and shape parameters for the twisted-stretched halos. | 2012.12284 | Twisted-Stretched2N.jpg |
0.445677 | 88770f61f7344f3284e1afb6dceec380 | The radial profile of Axes/r, angles and shape parameters for the twisted halos. | 2012.12284 | Twisted1N.jpg |
0.413147 | f8be39059c354fb3837de9bdd71a0c12 | The radial profile of Axes/r, angles and shape parameters for the twisted halos. | 2012.12284 | Twisted2N.jpg |
0.448208 | be73c11cbce044e5a7aa4613e6a17ea2 | The radial profile of the Axes/r ratio for two typical halos in our halo sample with (left) and without (right) including the substructures. It is seen that FOF group stars do not change the Axes/r ratio significantly. | 2012.12284 | axes_NEW2.jpg |
0.502648 | ac35bbea47b349e5ab378caeabeeb98c | (Left) The b-value vs the stellar halo mass for the full sample of galaxies with halo mass in the range (1-1.6) × 10^12 M_⊙ in TNG50. Yellow-stars refer to the distribution of the orbital circularity parameter for the MW-like galaxies in our sample. (Right) The distribution of the orbital circularity parameter, ε, for the disky and MW-like halos in our sample. Different halos are shaded using different colors. | 2012.12284 | bepsilon3.jpg |
0.439154 | ef819f6f93fd4f47a5f8a7e939a13266 | Comparison between the radial profile of the angles of different eigenvectors of DM and SH using the LSIM. See the caption of Figure <ref> for more details. | 2012.12284 | r-angle075.jpg |
0.443633 | f10eaf9ef9854403a1ad2126ae350c76 | Comparison between the radial profile of the angles of different eigenvectors of DM and SH using the EVIM. The eigenvectors are ordered as (min, inter, max), which are shown as (mi, in, ma) for brevity. There are in total 9 different angles. The DM and SH halos are more similar if the min-min, inter-inter and max-max angles are minimal and the rest of them are maximal. | 2012.12284 | r-angleN1.jpg |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.