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
|
---|---|---|---|---|
0.426362 | 5623077046c5403d851aa82da0cf89a2 | The energy integrated intensity of the resonance as measured in constant Q scans. The integrated intensity in absolute units was obtained by summing the data from 25 to 43 meV, and also by fitting a Lorentzian to data like Fig. <ref>. The latter method gives a resonance integral that reaches a maximum of ∼ 19 μ_B^2 at low temperature. There is a clear increase in the growth rate of the resonance intensity on entry to the superconducting state. Both methods show that a substantial fraction of resonant precursor is already present in the normal state. | cond | mat0308168-resonance_int.jpg |
0.422202 | 63cd26b669a1436099a4d91ce261b1a9 | The peak susceptibility at 12.4 meV and 24.8 meV as a function of temperature. For the normal phase, the solid curves are fits to the ω/T scaling analysis of Birgeneau et al. At 12.4 meV the normal state follows the scaling analysis. A clear suppression of scattering is observed in the superconducting state. The scattering at 24.8 meV in both the normal and superconducting states continues to grow on cooling and no longer follows the nearly constant temperature dependence predicted by ω/T scaling. The dashed line in the lower panel is a guide to the eye. The local and low-frequency susceptibility sensed by the NMR relaxation rate of ^63Cu is suppressed on cooling below a temperature T*, while the susceptibility at the antiferromagnetic wave vector increases uniformly until the onset of coherent superconductivity. NMR data is for YBCO_6.64 taken from Timusk and Statt. | cond | mat0308168-wT_compare.jpg |
0.444698 | 30fe960b90b1430a92367e5d135a0281 | A summary of the χ” is plotted for all normal state data below 16 meV energy transfer as a function of ω/T. The temperature dependence of the low energy scattering is very well described by ω/T scaling. | cond | mat0308168-wT_scale.jpg |
0.529896 | a550372f03ac4efdb055d1d0bc6c3fa7 | Logarithmic plots of the neutron elastic diffuse scattering intensity around a (100) Bragg peak, at different temperatures. | cond | mat0308170-fig1.jpg |
0.428299 | 5dfe962a7bfd4893bb01aee2cfa589ab | A smoothed logarithmic plot of the neutron elastic diffuse scattering intensity at 200 K. | cond | mat0308170-fig2.jpg |
0.491892 | 6338f43e64254a39b84c6df8556a8091 | Diffuse scattering intensities along the (110) direction near the (100) Bragg peak, measured at 10 K, 200 K, and 400 K. The intensity profiles are fitted using a broad Lorentzian function (solid lines), a narrow Gaussian (dashed lines) that describes the resolution-broadened Bragg peak, and a flat background. | cond | mat0308170-fig3.jpg |
0.481922 | c147120f77ca42e2a8040d8ca01d10d5 | Top frame: the correlation length ξ as a function of T. Bottom frame: integrated intensity I_0 as a function of T. | cond | mat0308170-fig4.jpg |
0.401003 | 6d31e82ffb5a4cc58483c5a7bfd59d0f | Plot of I_0/ξ^3 vs T. Parameters were obtained from the fits to the diffuse data. | cond | mat0308170-fig5.jpg |
0.453 | 59908c1837b84d41bb38f152439954b6 | Exponential fits (heavy dashed lines) to potential curves of Figure <ref> (light solid lines). () is mapped onto the background in gray to indicate the location of the phase field interface. | cond | mat0308173-potentialFits.jpg |
0.389675 | 0a843cb60b5a42938248a6d55b677033 | Potential-composition phase diagram for the parameters in Table <ref>, illustrating the bulk equilibrium between a electrode and a electrolyte with dissolved salt. Tie-lines denote different values of the quantity ( - ). The inset shows the position of this charge neutral phase diagram within the quaternary domain of the charged species. | cond | mat0308173-ternary.jpg |
0.458195 | c211707d12304ac8b52d2162a5b044d4 | The Hamiltonian of the transverse field Ising model. Each site is a spin-1/2 that interacts via Ising exchange with its nearest neighbors and can be flipped by the local x-magnetic field. | cond | mat0308176-Fig1.jpg |
0.425261 | 25deb2796479403ba21ea2c89a3f14c5 | A chain that terminates with the site 0 is dual to a chain that terminates with a site 0',which experiences no transverse field. The bond operator -h_00'1' is then the edge energy operator of the dual chain. | cond | mat0308176-Fig2.jpg |
0.49497 | 628078b84b7947fb827018c9647c9617 | Site decimation. Spin 0 is almost frozen in the x-direction due to the strong magnetic field h_0. Quantum fluctuations create a second nearest neighbor effective interaction between sites -1 and 1. This interaction is weaker than any of J_-10, J_01, h_0. | cond | mat0308176-Fig3.jpg |
0.490544 | a623cd735b914673822f0cdd4ac2a531 | Bond decimation. Sites 0 and 1 are frozen into one cluster by the strong Ising interaction, J_01. Quantum fluctuations produce an effective magnetic field, h̃_01=h_0 h_1/J_01, which flips the composite spin cluster. This field is weaker than any of h_0, h_1, J_01. | cond | mat0308176-Fig4.jpg |
0.531806 | 34a924626b864f4b8e2e0e10cd976e3d | Example of hierarchical decimation of a chain with four sites. First, site 1 is decimated. At a lower energy scale, sites 2 and 3 form a ferromagnetic cluster, which we denote (23). Cluster (23) then forms a cluster with site 0. The last process is a decimation of the cluster (023). The gound state wvae function of this chain is constructed perturbatively from the hierarchical wavefunction , which is given in Eq. (<ref>). | cond | mat0308176-Fig5.jpg |
0.556316 | 730f1be2fb8f4120a28ba45641d604c2 | Evolution of σ^x_0 in a bond decimation. When the end site forms a cluster with its neighbor, the operator σ^x_0 gets renormalized and gains a factor of h_1/J_01. | cond | mat0308176-Fig6.jpg |
0.464847 | d8d0a1a5bfaa4888a4a951767af28d3b | Evolution of σ^x_0 in a site decimation. Although σ_0^x obtains an expectation value, its fluctuations are still influenced by the state of site 1. This is reflected in the renormalization of σ_0^x as J_01^2/h_0^2σ_1^x which gives rise factor of J_01^2h_0/h_1^3 in the correlation functions of 0. | cond | mat0308176-Fig7.jpg |
0.440278 | c8cd28fe02cc420eb2d765457a2ae2ce | Progress of alloy electrodeposition upon step from = -102 to = -107. () is mapped onto the background in gray to indicate the location of the phase field interface. (Color) | cond | mat0308179-alloy.jpg |
0.411211 | ba003e9162c54c0fab5a6d23d37189f3 | Interface profiles for steady state electrodeposition with = -102. The concentration profiles for + and - are almost coincident on this scale. () is mapped onto the background in gray to indicate the location of the phase field interface. | cond | mat0308179-alloy0A.jpg |
0.52943 | d9a11cd21ab74a5e99da4956237b7bfa | Concentration + of + in the deposited electrode as a function of overpotential . | cond | mat0308179-alloyConcentration.jpg |
0.473219 | b637b52989c2454c8db46915d6f7d5f6 | Potential-composition phase diagram for the parameters in Table <ref>, illustrating the bulk equilibrium between a electrode and an electrolyte containing salt dissolved in . Tie-lines denote different values of the quantity ( - ). The inset shows the position of this charge neutral phase diagram within the quaternary domain of the charged species. | cond | mat0308179-quaternary.jpg |
0.455447 | 7fbc9ace6adc4795bc70e78713a7d093 | Schematics of the inhomogeneous dopant distribution model proposed in this paper. x_inner and x_outer are the dopant concentrations for the inner and outer shells respectively. n_inner and n_outer are the number of cation sites enclosed within the inner and outer shells, respectively. Black circles show a random configuration of dopant atoms. For the low dopant concentration (“HCl washed”) samples, the restriction of doping to a small inner shell increases the probability of pairs as compared to the case where the dopant atoms are uniformly distributed over the entire NC, hence reducing M_5T/M_sat. For high dopant concentration (“as prepared”) samples, x_outer > x_inner reduces the probability of pairing between surface and interior dopant atoms as compared with the uniformly distributed case, hence increasing M_5T/M_sat . | cond | mat0308180-dope.jpg |
0.437449 | ad06595bb72e4a2bab7914802ac223c5 | The baby skyrmion with m=1 in Skyrme theory, where we have put α/μ^2=1 for simplicity. Notice that ϱ is in the unit of √(α)/μ. | cond | mat0308182-babysk.jpg |
0.48803 | 15c109f3644749d891b8b61c5eb49002 | The non-Abelian vortex with m=1 in gauge theory of two-component BEC. Here we have put g=√(λ)=1, and ϱ is in the unit of κ. | cond | mat0308182-bec.jpg |
0.575521 | 3d19d00d6cc045d884c6f814bae77d3e | The non-Abelian vortex with m=1 in two-gap superconductor which has 4π/g flux. Here we have put g=1 and √(λ)=2, and ϱ is in the unit of ρ_0. | cond | mat0308182-scv.jpg |
0.530439 | 66c437bc7293439d9082cf07d0568a40 | The non-Abelian vortex with m=1 in two-gap superconductor which has 2π/g flux. Notice that here we have ρ(0) 0. | cond | mat0308182-scv1.jpg |
0.448723 | 5c4a03562e2947bd8ad885a9d8ca92d1 | The non-Abelian vortex in the Gross-Pitaevskii theory of two-component BEC. Here we have put n=1, and ϱ is in the unit of κ. Dashed and solid lines correspond to δμ/μ= 0.1 and 0.2 respectively. | cond | mat0308182-twobec.jpg |
0.434677 | eafcae450ac343bbac2ca8e1e0cab1c3 | Left: Photoemission spectrum (PES) for the half-filled 5× 4 cluster, J/t=0.5., t'/t=-0.2, t”/t=0.1. Right: Dispersion of the quasiparticle peak as extracted from the numerical spectra (top) compared to the theoretical single hole dispersion ϵ_1( k) in the spin-Peierls phase (bottom). | cond | mat0308184-fig1.jpg |
0.46866 | 7976a18118b54f66aeaaf04de981cc20 | Single particle spectrum for the 5× 4 cluster with two holes. Parameter values are J/t=0.5., t'=t”=0 (Left) and J/t=0.5., t'/t=-0.2, t”/t=0.1 (Right). Only the parts near the Fermi energy E_F/t≈ 1.5 are shown. | cond | mat0308184-fig2.jpg |
0.414577 | bd43d5845ba74a39b88be4c197872cd3 | Left: Single-particle spectral function from bond-operator theory for t'/t=-0.1, t”/t=0.05, the Fermi energy (verical line) corresponds to a hole density of 0.16. Right: Single hole dispersion (top) and momentum distribution n( k) (bottom) for 5× 4 and the same parameters. | cond | mat0308184-fig3.jpg |
0.430334 | df79ef78036c4cf0926fe5ee7d1bcf04 | A 2D model of the epitaxial film and substrate showing the particle configurations in the coherent state. The two layers at bottom are held fixed while all other are free to move. Filled circles represent the the epitaxial film and open circles the substrate. | cond | mat0308186-fig1.jpg |
0.444952 | 1b72013075a14b66927f595985710577 | Minimal Energy path for compressive strain f=+8% as a plot of energy barrier Δ E vs reaction coordinate S. Snapshots configurations (a), (b) and (c) correspond to the labels in the energy profile (top right). Closed line in (c) is the Burgers circuit around the dislocation core. The energy barrier is in units of interatomic potential strength ϵ and the reaction coordinate S is in units of equilibrium distance r_ ss. | cond | mat0308186-fig2.jpg |
0.426766 | 118603d07d5c4dcf967ab03c6821a007 | Minimum Energy path for tensile strain f=-8% as a plot of energy barrier Δ E vs reaction coordinate S. Snapshots configurations (a), (b) and (c) correspond to the labels in the energy profile (top right). Closed line in (c) is the Burgers circuit around the dislocation core. The energy barrier is in units of interatomic strength ϵ and the reaction coordinate S in units of equilibrium distance r_ ss. | cond | mat0308186-fig3.jpg |
0.417716 | 56a5289a46fe4078b79b00dcbdfd4685 | Saddle point configurations for different mechanisms of stress relaxation (a) glide mechanism for tensile strain (b) glide mechanism for compressive strain (c) climb mechanism for compressive strain. Filled circles represent the the epitaxial film and open circles the substrate. | cond | mat0308186-fig4.jpg |
0.50191 | 1f07f46cb4004cd1869e2ff4eb39aa0f | Energy barrier Δ E (in units of ϵ) as a function of film thickness (number of layers) for different misfit values. Squares symbols correspond to f=± 4 %, stars to f=± 5%, triangles to f=± 6%, and circles to f=± 8%. Solid and dotted lines correspond to compressive f > 0 and tensile f < 0 strains, respectively. | cond | mat0308186-fig5.jpg |
0.49326 | a9ca83dd7f904755ae48875a8f5e50aa | Energy barrier Δ E (in units of ϵ) as a function of film thickness (number of layers) at misfit 5%, for the 5-8 (squares) potential and 6-12 (circles) potential (cut off 1.5 r_ ss). Solid and dotted lines correspond to compressive f > 0 and tensile f < 0 strains, respectively. Here the system size is L=20. | cond | mat0308186-fig6.jpg |
0.438726 | 785660a03f77488ab6fe4cc0594e8b8f | Energy profile of the minimum energy path for (a) compressive and (b) tensile strain and for different misfits. Energy in units of ϵ and S in units of equilibrium distance r_ ss. | cond | mat0308186-fig7.jpg |
0.488629 | fff5a01d85b1459e872c7dab55e051e5 | Energy barrier Δ E (in units of ϵ) as a function of film thickness (number of layers) for smaller sample size (20 atoms per layer) and different misfit values for the 5-8 potential: f=± 5% ( stars), and f=± 8% (circles). Solid and dotted lines correspond to compressive f > 0 and tensile f < 0 strains, respectively. | cond | mat0308186-fig8.jpg |
0.420126 | 7b0c1822fd8f41e98eaf152ed52c7012 | Energy barrier Δ E (in units of ϵ) as a function of film thickness (number of layers) for different misfit values for long ranged 5-8 potential (cut off 2.1 r_ ss): f=± 4 %(squares), f=± 5% (stars), f=± 6% (triangles), and f=± 8% (circles). Solid and dotted lines correspond to compressive f> 0 and tensile f < 0 strains, respectively. | cond | mat0308186-fig9.jpg |
0.684586 | 045f6c7bbb6e4078bac4e5ea3ce8e71b | An adherent cell actively pulls on its soft environment through cell-matrix contacts. Experimentally, one finds that cells orient themselves in the direction of maximal stiffness of the environment. With this cartoon, we present one possible mechanism by which active mechanosensing in an elastically anisotropic medium might lead to cell orientation. The local elastic environment is represented by linear springs with different spring constants K, as indicated by differently sized springs. For upregulation of a contact, the cell has to invest the work F^2/2K. Therefore, upregulation is more efficient for larger K. (a) In an isotropic environment, all spring constants are the same, growth at different contacts is similar and the cell does not orient. (b) If spring constants are largest in one specific direction, the corresponding contacts outgrow the others and the cell orients in the direction of maximal stiffness of the environment. In this paper, we use the cellular preference for large effective stiffness and modelling of the extracellular environment by linear elasticity theory to predict cell positioning and orientation in soft media. | cond | mat0308187-Fig1.jpg |
0.44981 | aa881ec333f541e8bd174e9e26b73cb5 | Adjusting cell position and orientation in such a way that the cell senses maximal effective stiffness in its environment is equivalent to minimizing the quantity W, the amount of work the cell invests into the elastic surroundings in the presence of external strain. In the presence of mechanical activity, sample boundaries induce external strain which can result in different cell organization. (a) Δ W for a cell with dipole strength P which is a distance d away from the surface of an elastic halfspace with rigidity E, plotted in units of P^2/E d^3 as a function of angle θ between cell orientation and surface normal (rescaled by 256 π). Solid and dashed lines correspond to Poisson ratios ν = 1/2 and ν = 0, respectively. Irrespective of ν, the optimal orientations (minimal Δ W) are perpendicular (θ = 0) and parallel (θ = π/2) to the surface for clamped and free boundaries, respectively. Since |Δ W| increases if d decreases, the overall mechanical activity of a cell increases towards a clamped surface (Δ W < 0), but decreases towards a free surface (Δ W > 0). (b) Δ W for a cell in an elastic sphere of radius R, plotted in units of P/E R^3 as a function of distance r to the sphere center in units of R for ν = 1/3 (rescaled by 15/8). Solid and dashed lines are parallel (θ = π/2) and perpendicular (θ = 0) orientations, respectively (all other orientations yields curves which lie inbetween the ones shown). Like in an elastic halfspace, parallel and perpendicular orientations are favored (minimal Δ W) for free and clamped boundaries, respectively. For clamped boundaries, mechanical activity is favored (smaller Δ W) towards the surface. For free boundaries, mechanical activity is disfavored (larger Δ W) towards the surface. | cond | mat0308187-Fig2.jpg |
0.463013 | 15ebb7559add456ab75fa56b4396d49f | Predicted cell orientation in a hydrogel close to a surface (a,b) and on elastic substrates (c,d). (a) Cells prefer the direction of maximal effective stiffness. Thus, they orient perpendicular to a clamped surface. (b) For a free surface, this direction is parallel to the surface. (c) Cells close to a boundary between soft (left) and rigid (right) regions prefer analogous orientations as cells close to clamped and free surfaces in a hydrogel, respectively. (d) Cells interact elastically to form strings, because in nose-to-tail alignment, the mechanical activity of one cell triggers the one of the other cell, thereby forming a positive feedback loop. | cond | mat0308187-Fig3.jpg |
0.448305 | 2d6ac928a4c847ebbd156e15c5a9415c | Monte Carlo simulations of elastically interacting cells in an external strain field. The temperature used in the simulation represents the stochastic element of the process of cell organization. Without external strain, cells form strings. In its presence, strings align in parallel. | cond | mat0308187-Fig4.jpg |
0.469018 | 48d15c6d36084fd6bd329d3636e494df | The derivation of Möbius interionic potentials. Three types of virtual structure are constructed: B3, T1, and B2, for the purpose of extracting one from the totally three short-range interactions: cation-anion, anion-anion, and cation-cation. Of each energy curve, the long-range Coulomb energy has been pre-subtracted from the total energy. | cond | mat0308189-fig1.jpg |
0.46198 | 6a16de3089a1427cb9415fb4096d4e32 | The 0-K Gibbs free energy surfaces of R3̄m symmetry at (a) zero, (b) P_tr and (c) 30GPa, respectively. | cond | mat0308189-fig2.jpg |
0.460261 | 8e52f3bccb784941bc84839e46eb3687 | The Gibbs free energies of relaxed B1 and B2-structure RbCl crystals increase with external pressure. The two curves have a crosspoint at P=1.09GPa. | cond | mat0308189-fig3.jpg |
0.452224 | c070a58610c44ef2b6b57e287631e21e | The thermal expansion of perfect B1-RbCl crystal calculated with molecular dynamics. Temperature varies from 100K to 900K. The linear coefficient of thermal expansion is defined as (1/L)(dL/dT), where L is the lattice constant, and the derivative is calculated as a central derivative with a temperature step of 200K. | cond | mat0308189-fig4.jpg |
0.45241 | f1e6af9f933649398c82ae6032981d64 | The volume-temperature curves of the coexist-phase lattice and the perfect lattice. The abrupt change of volume with the one-phase method occurs at a temperature about 260K higher than that of the two-phase method. | cond | mat0308189-fig5.jpg |
0.527995 | 9bf5ef4a9d754749a64e473d1022f88a | The initial configuration of a coexist-phase lattice (a) and possible final configurations (b)(980K), (c)(1000K) after equilibrating the system at different temperatures. Dark circles refer to ions that were initially in the solid phase; gray circles refer to ions initially in the liquid phase. The system has almost completely solidified as (b); and has almost completely melted as (c). | cond | mat0308189-fig6.jpg |
0.512863 | fc2bddc352644331b6da0dc54cfdc891 | The radial distribution function (RDF) of the final configurations at different temperatures. The abrupt change of RDF indicates the occurrence of melting. (a) shows the RDF of cation-cation and (b) shows that of cation-anion. | cond | mat0308189-fig7.jpg |
0.52676 | d28741436824437da37f6681e3d87e55 | The mean-square displacement (MSD) of all ions as the function of time at different temperatures. The change of MSD curves indicates the occurrence of melting: at low temperatures the MSD oscillates near its balance, while it increases with time at temperatures above melt-point. | cond | mat0308189-fig8.jpg |
0.496883 | fd09a2d531a5404fbb30944372705e40 | Table II. | cond | mat0308189-table2.jpg |
0.444724 | a9bedd7c0f5e47b0b870823c3ce054e4 | Table III. | cond | mat0308189-table3.jpg |
0.432149 | df3657d9792e4de6858f014671ce76a1 | Scanning microscope image of the two coupled qubits surrounded by the DC-SQUID. A part of the microwave coupler is visible on the right. The width of the qubits is 5 μm. | cond | mat0308192-Figure1.jpg |
0.473282 | b6b8735049204f579e2be7d3451a0ee3 | Energy levels of two identical coupled qubits as a function of the externally applied flux. Far away from the degeneracy point |h|≫ t, the energy states resemble classical states. The coupling j shifts the energy of the ferromagnetic states , up, while the energy of the antiferromagnetic states , is lowered. In the vicinity of the degeneracy point |h|≲ t, the classical states mix due to the tunnel coupling t. Due to the symmetry of the problem, the anti-symmetric energy state - does not mix with symmetric states. | cond | mat0308192-Figure2.jpg |
0.464021 | dd39286ea8684a8190fccf0d8cec821c | (a) Energy levels of two different coupled qubits as a function of the externally applied flux. Due to the difference between the two qubits, the degeneracy of the two antiferromagnetic states , is lifted (b) Transition matrix elements |T_if|=|<Ψ_f|σ_1^z+σ_2^z|Ψ_i>| for the transitions from the ground state to the three excited states. The matrix element for the transition from the ground state to the third excited state is very small except for the vicinity of the degeneracy point. The parameter values are taken from the values obtained by fitting the experimental data (Fig. <ref>). | cond | mat0308192-Figure3.jpg |
0.467511 | 92dff98984b749ceab27be994cfd7600 | Spectroscopy measurements. (a) SQUID signal, i.e. magnetization of the two qubits, as a function of the applied magnetic field. The microwaves induce transition from the ground state to the first and second excited state, which appear as dips and peaks in the magnetization. The outer peaks and dips are due to the first excited state and the inner due to the second. (b) Microwave frequency versus peak and dip position. In the vicinity of the degeneracy point (Φ≈ 1/2Φ_0), the resonances deviate from straight lines because the energy levels are formed by superposition of the classical states. The straight line is a fit of the two qubit Hamiltonian (Eq. <ref>) to the data. The inset compares the data obtained in the vicinity of 3/2Φ_0 with the data obtained at 1/2Φ_0. The shift is due to the influence of the different qubit areas which is three times higher. | cond | mat0308192-Figure4.jpg |
0.446177 | ae103dde40c141d7a5841dd24cca7d19 | Body-centered cubic lattice (black points) obtained by applying an infinite-strength external potential at the white points of a simple cubic lattice. Hard cubes with σ=2 can only be placed at black points, thus the interaction potential becomes nearest-neighbor exclusion in the bcc lattice. | cond | mat0308195-bcc.jpg |
0.467504 | 03184896d9104b96b28a6d51ffd7fad3 | Hard cubes with edge length σ=2 in a simple cubic lattice. Their centers of mass are constraint to lie on the plane s_1+s_2+s_3=0. Filled circles mark the sites excluded by the cubes which lie on this plane. Notice that on the plane, cubes behave as the hard-hexagon lattice gas with nearest-neighbor exclusion. The diagram at the lower right corner helps to visualize the relative position of the cubes in the figure. | cond | mat0308195-hexagonos2.jpg |
0.47634 | 25e546d3ff9b4c568c893b01f093597b | Hard cubes with edge length σ=2 in a simple cubic lattice. Their centers of mass are constraint to lie on the plane s_1+s_2+2s_3=0. Filled circles mark the sites excluded by the cubes which lie on this plane. Notice that on the plane, cubes behave as the hard-square lattice gas with nearest-neighbor exclusion. The diagram at the lower right corner helps to visualize the relative position of the cubes in the figure. | cond | mat0308195-rombos2.jpg |
0.466597 | ae94abbd5ed747359fba07fc0b459814 | Zero-field μSR spectra in measured across the superconducting transition temperature T_c1=1.82 K. The corrected asymmetry was obtained by subtracting the temperature-independent contribution from the silver plate (0.057). The curves are fits to the model I given by Eq. (1). The relaxation rate becomes stronger below T_c1. | cond | mat0308196-fig1_aoki.jpg |
0.420401 | 7fa4af3d0ac24b0db86ecb2b9433acac | Temperature dependences of the width of the internal field distribution Δ, the relaxation rate Λ in zero-field (ZF) and in μ_0 H_ LF=0.01 T, and the specific heat divided by temperature C/T. The vertical lines indicate two consecutive transitions at T_c1 and T_c2 determined from the C/T data. | cond | mat0308196-fig2_aoki.jpg |
0.462751 | ef029696716a4aa488ec5ca7585ade9c | Longitudinal field H_ LF dependence of the relaxation rate Λ. The curves are fits to the model given by Eq. (3). | cond | mat0308196-fig3_aoki.jpg |
0.50642 | 51701676602d4b4c89557a9a4d35cc48 | Poly-dT_nu F(t), for several values of λ, with: N=30, d=12, x=N+d/2, T=2^oC, V/V_C=2, ω_B=10^2Hz, ω=1, and z≈ 1/2. The left peak represents the non-translocated events, whereas the other two peaks represent translocation. Inset: The range for which λ yields three peaked F(t) is shown to be 0.1≤λ≤0.3, when given the above parameters. | cond | mat0308199-f1cm.jpg |
0.484503 | e7b07509abed4c3ba426317b492509aa | Poly-dT_nu F(t), for several values of ω_A and fixed ω_B (ω_B=10^2Hz), with λ=1/4, and the other parameters as in Fig. 1. Inset: For small values of ω, ω≲ 10^-2, F(t) displays one translocation peak that corresponds to conformation A, whereas for large values of ω, ω≳ 10^2, F(t) displays one translocation peak that corresponds to conformation B. For ω≈ 1 two translocation peaks are obtained. | cond | mat0308199-f2cm.jpg |
0.389263 | 3243c9a7b8944cdab385aacb7fc4b0cb | a: t_m,i for poly-dT_nu, as a function of V/V_C for the same parameters as in Fig. 1 and V_λ=350mV. t_m,1 is almost independent of V_C/V in contrast to the pronounced dependence of t_m,2 and t_m,3. b-c: t_m,2^-1 and t_m,3^-1 depend linearly and quadratically on V/V_C, respectively. The solid lines through the circles are polynomial fits. | cond | mat0308199-f3.jpg |
0.41812 | 615d3150625a435fafe3803f7b2ea6b9 | t_m,2^-1 for poly-dT_mu for the same parameters as in Fig. 3 except for V_λ=120mV. This value for V_λ leads to 0.625≤λ(V)≤ 0.875 and accordingly for one translocation peak. The solid line is a polynomial fit. Inset: t_m,1 and t_m,2 behave qualitatively the same as for the case V_λ=350mV. | cond | mat0308199-f4.jpg |
0.48948 | 74006edf15624dbfb2393226ed52b9bb | G(t) for poly-dA_nu, full curve, for V=0, and the initial state x=n/2, ω_A=10^-2Hz, ω=1/2 and the other parameters as in Fig. 1. Also shown is the approximate cdf G_ap(t), dashed curve, and its modified version, dotted curve. Inset: F(t) for poly-dA_nu for the corresponding cdf shown in the main Fig. | cond | mat0308199-f5.jpg |
0.40735 | 5af55df20b234cd89faf05bdeeda3e2e | a) Sheet resistance(Ω/SQ) as function of temperature (T) for samples A - D. b) Ω/SQ as function of T for Co, Cr, Mn, V implanted p^+GaAs:C with 3 × 10^16 cm^ - 2 dose. c) AC magneto-transport measurement of sample C at 50 K with excitation current of 100 μA. d) AC Hall Effect measurement of samples A - C at 70 K. | cond | mat0308200-fig1.jpg |
0.412033 | a6d411100cea456094441fb842a66f25 | a) M_R as a function of temperature (T) for Mn implanted p^+GaAs:C (near T_C for Sample C, inset). b) Magnetization (M) as a function of T for sample X and Y with Curie-Weiss Law fit. M as a function of applied field at 5 K for sample C (c) and sample X and Y, offset for clarity (d). | cond | mat0308200-fig2.jpg |
0.390245 | 1734623520694963b89745c53fccae5a | a) M_R and AC resistivity (I = 100 μA) as a function of temperature for sample C (T_C∼ 280 K) at H = 0. Anomalies in transport properties correspond to magnetic properties suggesting changes in resistivity are due to magnetic ordering in the sample. b) AC-Hall measurement at various temperature for sample C. Inset shows low field and high field fit to the Hall response corresponding to ordinary and extraordinary Hall coefficient for Sample C, indicating non-linear Hall response below ∼275 K. | cond | mat0308200-fig3.jpg |
0.452687 | c2e766b67c6a46e9bff4d757f5d1c492 | Magnetic-field dependence of ρ_yx for a YBCO crystal with y=6.95 (optimum doping) at various temperatures. | cond | mat0308201-Segawa_Fig1.jpg |
0.477141 | ab2aa79b57fb4453931f9a9c6eff25f8 | y dependences of the Hall mobility μ_ H at 125, 150, 200 and 300 K with the current along the b-axis (a) and the a-axis (b). | cond | mat0308201-Segawa_Fig10.jpg |
0.450394 | 00af85cdec2b489894e5d0b9d975892c | The temperature dependences of (a) ρ_a and ρ_b, (b) R_ H^ meas and R_ H^ pl(iso) for Y124 and YBCO samples with y=7.00. (c) T dependences of σ_xy and (d) Θ_ H(b)^ pl vs. T^2 plot for Y124 and YBCO samples with y=7.00 and 6.75. | cond | mat0308201-Segawa_Fig11.jpg |
0.504914 | b9eefd214ff94763ae6103c45cfa2ffa | R_ H(T) for two measurements on the same crystal at y=6.95 (optimum doping), with j∥ a(open squares) and j∥ b (solid circles). | cond | mat0308201-Segawa_Fig2.jpg |
0.485508 | ddd9e3110cdd4855ba6d8f5fdec0ec73 | (a) T dependences of σ_xy for samples with 5 representative y values. (b) y dependences of σ_xy at 125, 150, 200 and 300 K. | cond | mat0308201-Segawa_Fig7.jpg |
0.391004 | 73065e20c0ba4c57984f648fcbf3c7a9 | (a) Θ_ H(b)^ pl vs T^2 plot for y=6.30–7.00, where dashed lines are fits to the data by AT^2 and solid lines are fits to the data by AT^α+C. (b, c, d) y dependences of the fitting parameters A, α and C, respectively. | cond | mat0308201-Segawa_Fig8.jpg |
0.403522 | c8dc8da9fd1c49beaa620690478e4469 | (a) Θ_ H(a) vs T^2 plot for y=6.30–7.00, where solid lines are fits to the data by AT^α+C. (b, c, d) y dependences of the fitting parameters A, α and C, respectively. | cond | mat0308201-Segawa_Fig9.jpg |
0.478788 | cab03ff956a74bc48961771a143ed227 | Sliding nanotubes. | cond | mat0308206-fig1.jpg |
0.479002 | 8c98b4ed09054e81b6950e0379889f02 | Probability distribution, Π(p_τ) of the binned power in time bins τ=0.5, 1, 2, 4, 8 and 16 ms. The distributions are displaced vertically for clarity. The average power for all distributions is p̅_τ=1, indicated by the dashed vertical line. | cond | mat0308212-Hist_PRL.jpg |
0.427207 | fab5e756a3b149c19e8a1bb75d55630f | (A) ln[Π(p_τ)/Π(-p_τ)] versus p_τP for τ ranging from 0.5 to 16 ms. (B) ln[Π(p_τ)/Π(-p_τ)]/τ versus p_τP̅ (P=356 m^2s^-3). The solid line shows the slope of the collapsed curves. A dashed line of slope 1/T_gran is drawn for comparison. | cond | mat0308212-Power_ln_PRL2.jpg |
0.456556 | 648d6f542ab04736b222b8081ff25ba3 | A sample (∼0.6%) of the normalized power trace for the subsystem in consideration. The horizontal line shows the average power, p̅=1. | cond | mat0308212-Power_trace_PRL.jpg |
0.430619 | 206358b1001746e7a7640f8c297f0fd2 | T_eff (left axis), defined as the inverse of the slope in a graph such as Figure 4B, and T_gran (right axis), the granular temperature (1/2⟨ v^2⟩), as a function of area fraction. T_eff is numerically larger but follows a similar trend. | cond | mat0308212-Teff_PRL.jpg |
0.46021 | 12447587fd4941638846bd59fb1805c7 | Sketch of the experimental cell. The dashed rectangle is a window measuring 10dx21d, fixed in the laboratory frame, in which we study the flux of kinetic energy. | cond | mat0308212-cartoon_PRL.jpg |
0.440577 | b7f49efce6424ed9976dc1e704fcc1c6 | 1D Bose gas in an intermediate regime: T_rel=10 and γ=10. Left panel shows the momentum distribution, while the right panel is the second-order spatial correlation function. Here, the distance x is in units of the healing length ξ, which can be expressed as ξ=Λ_th√(2T_rel/2γ). Solid lines are the results of numerical simulations with 738200 stochastic trajectories. Dashed lines are the results for γ=0 ideal Bose gas (ID), γ→∞ Tonks-Girardeau (TG) gas, and T_rel→∞ Boltzmann (BL) gas, shown for comparison. | cond | mat0308219-g10t10structxi.jpg |
0.497304 | cebfc0f18ba64a67a929eb7d8b4c9b9a | Comparison of numerical calculation (solid lines) to exact results <cit.> (asterisks), as well as to Ideal (ID) and Tonks-Girardeau (TG) gases at the same temperature T and chemical potential. Left panel: particle number density. Right panel: energy per particle ℰ. | cond | mat0308219-g10t10yang.jpg |
0.451741 | adf33e36088f4458960648c30fb2e51f | Schematic showing the spin-LED geometries for edge emission (a) and vertical emission (b). Shape anisotropy tends to force situation (a) without a large vertical magnetic field or nanofabrication. | cond | mat0308220-figure1.jpg |
0.493449 | dca57bdba13d44089b9da257e2577603 | (a) Total edge emission intensity, I_+ + I_- (thicker solid line), the intensity difference between right-circularly and left-circularly polarized emission, I_+ - I_- (dashed line) and luminescence polarization P_ℓ (dotted line) for heavy-hole spin injection into a 100 Å In_0.13Ga_0.87As QW with hole density 10^15 cm^-3 and P_d,s = 52% (Δμ^h =0.57 meV). (b) energy-integrated edge-emission polarization (solid line) and P_d (dashed line) versus Δμ^h. (c) same as (b) but for light-hole spin injection. The temperature is 6K. A 3 meV Gaussian linewidth smoothes the luminescence spectrum. | cond | mat0308220-figure2.jpg |
0.407292 | 63e4ecae92644d6f8b56438e4def64d7 | Vertical emission from the system in Fig. <ref>(ab) with the same parameters except P_d,s = 27% (Δμ^h =0.28 meV). | cond | mat0308220-figure3.jpg |
0.444577 | 7d6c723489d4451fac86767c56b0bd07 | Maximal P_ℓ (solid line) and P_d (dashed line) in the QW as a function of hole density for in-plane emission at T=6K. The inset illustrates the total electroluminescence (solid line) and P_ℓ with hole density 10^16 cm^-3. | cond | mat0308220-figure4.jpg |
0.394144 | 31febfe3ea384201a0934a384773ad05 | Luminescence and density polarizations versus the Fermi-level splitting for spin-LED structures using spin-polarized electrons. Panels (a) and (b) are for edge emission and vertical emission, respectively. The solid lines describe the luminescence polarization and the dashed lines describe the density polarization. The electron density is 10^15 cm^-3 and the temperature is 6 K. | cond | mat0308220-figure5.jpg |
0.453342 | 4f055550372048cf93f66bb85449bc38 | (a) AFM micrograph of the ring structure. Oxide lines (bright lines) fabricated by AFM lithography lead to insulating barriers in the two-dimensional electron gas. Areas marked pg1-pg4 are used as lateral gates to tune the conductance of the four arms of the ring and eventually act as plunger gates for the dots. (b) Schematic arrangement of the ring with its four terminals (c) Non-local voltage detected in the open regime of the ring. (d) Conductance through a quantum dot induced in segment 1 as a function of plunger gate voltage. Here, segment 3 was completely pinched off. | cond | mat0308223-fig1.jpg |
0.442402 | 5febea5154d44ec3a3f79b50d7adf659 | Color plots of (a) the non-local voltage V_nl, (b) the current I_ and (c) I_ as a function of the two plunger gate voltages V_pg1 and V_pg2 which tune the quantum dots in segments 1 and 2 of the ring. | cond | mat0308223-fig2.jpg |
0.455163 | 28f4f14b19a94b568a0a959bac00eb16 | AB oscillations in non-local voltage V_nl as function of the plunger gate voltages tuned along trace `linear2'. (a) The currents I_ and I_ at a magnetic field of 95 mT. (b) Color plot of V_nl as function of plunger gates and magnetic field. (c) Cross-sections along dashed lines in (b) giving V_nl vs. magnetic field. The dashed lines mark constant values of magnetic field where a π phase shift of the AB oscillations can be seen, as gate voltages are tuned across the conductance peak in dot 2. | cond | mat0308223-fig3.jpg |
0.397191 | 9b6c6ffe5747476eb732d4569b24aba6 | AB oscillations along the different traces indicated in Fig. <ref>. (a) sweep `linear 2' (b) sweep `diag 1' (c) sweep `linear 1' and (d) sweep `diag2'. Vertical lines are guides to the eye. Dashed horizontal lines indicate, where maxima in the magnetic field averaged contribution to V_nl occur. | cond | mat0308223-fig4.jpg |
0.454367 | ea68e18af8c24788bd2cfc5593ef717d | Pbnm orthorhombic crystal structure of YVO_3. | cond | mat0308224-Figure1.jpg |
0.467123 | d6b0226bcb11426c8b8215b2c7b18bbb | Vanadium 3d partial density of states for spin-up and spin-down electrons, obtained in local spin-density approximation with the Hubbard potential U taken into account. N-1 and N+1 correspond to the electron removal and electron addition states, respectively. | cond | mat0308224-Figure10.jpg |
0.489218 | b172660222cf4702a7c3803ca6f9a8eb | The dielectric functions from Fig. <ref>, the real and imaginary parts, with subtracted 250 K data. The fit curve shows a simultaneous fit to the real and imaginary part for the 80 K data. | cond | mat0308224-Figure11.jpg |
0.402136 | 2720bf01cd9147069adc4520947e2a3e | Temperature dependence of the bands position (left) and their optical strength together with the total spectral weight of the three bands (right). | cond | mat0308224-Figure12.jpg |
0.480593 | 8a0a4dc31f4c4293826992b8d50199a9 | (a) Temperature dependence of the Kerr rotation measured in the polar geometry with H||a. (b) The Kerr rotation Θ_K together with the Kerr ellipticity ε_K in the low temperature phase. (Insert) Enlarged view of the resonances at 1.8 and 2.2 eV. | cond | mat0308224-Figure13.jpg |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.