Sample records for observed velocity gradients

  1. Radial and latitudinal gradients in the solar internal angular velocity

    NASA Technical Reports Server (NTRS)

    Rhodes, Edward J., Jr.; Cacciani, Alessandro; Korzennik, Sylvain G.; Tomczyk, Steven; Ulrich, Roger K.; Woodard, Martin F.

    1988-01-01

    The frequency splittings of intermediate-degree (3 to 170 deg) p-mode oscillations obtained from a 16-day subset of observations were analyzed. Results show evidence for both radial and latitudinal gradients in the solar internal angular velocity. From 0.6 to 0.95 solar radii, the solar internal angular velocity increases systematically from 440 to 463 nHz, corresponding to a positive radial gradient of 66 nHz/solar radius for that portion of the solar interior. Analysis also indicates that the latitudinal differential rotation gradient which is seen at the solar surface persists throughout the convection zone, although there are indications that the differential rotation might disappear entirely below the base of the convection zone. The analysis was extended to include comparisons with additional observational studies and between earlier results and the results of additional inversions of several of the observational datasets. All the comparisons reinforce conclusions regarding the existence of radial and latitudinal gradients in the internal angular velocity.

  2. The lateral variation of P n velocity gradient under Eurasia

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Yang, Xiaoning

    We report that mantle lid P wave velocity gradient, or P n velocity gradient, reflects the depth and lateral variations of thermal and rheological state of the uppermost mantle. Mapping the P n velocity gradient and its lateral variation helps us gain insight into the temperature, composition, and dynamics of the uppermost mantle. In addition, because P n velocity gradient has profound influence on P n propagation behavior, an accurate mapping of P n velocity gradient also improves the modeling and prediction of P n travel times and amplitudes. In this study, I used measured P n travel times tomore » derive path-specific P n velocity gradients. I then inverted these velocity gradients for two-dimensional (2-D) P n velocity-gradient models for Eurasia based on the assumption that a path-specific Pn velocity gradient is the mean of laterally varying P n velocity gradients along the P n path. Result from a Monte Carlo simulation indicates that the assumption is appropriate. The 2-D velocity-gradient models show that most of Eurasia has positive velocity gradients. High velocity gradients exist mainly in tectonically active regions. Most tectonically stable regions show low and more uniform velocity gradients. In conclusion, strong velocity-gradient variations occur largely along convergent plate boundaries, particularly under overriding plates.« less

  3. The lateral variation of P n velocity gradient under Eurasia

    DOE PAGES

    Yang, Xiaoning

    2017-05-03

    We report that mantle lid P wave velocity gradient, or P n velocity gradient, reflects the depth and lateral variations of thermal and rheological state of the uppermost mantle. Mapping the P n velocity gradient and its lateral variation helps us gain insight into the temperature, composition, and dynamics of the uppermost mantle. In addition, because P n velocity gradient has profound influence on P n propagation behavior, an accurate mapping of P n velocity gradient also improves the modeling and prediction of P n travel times and amplitudes. In this study, I used measured P n travel times tomore » derive path-specific P n velocity gradients. I then inverted these velocity gradients for two-dimensional (2-D) P n velocity-gradient models for Eurasia based on the assumption that a path-specific Pn velocity gradient is the mean of laterally varying P n velocity gradients along the P n path. Result from a Monte Carlo simulation indicates that the assumption is appropriate. The 2-D velocity-gradient models show that most of Eurasia has positive velocity gradients. High velocity gradients exist mainly in tectonically active regions. Most tectonically stable regions show low and more uniform velocity gradients. In conclusion, strong velocity-gradient variations occur largely along convergent plate boundaries, particularly under overriding plates.« less

  4. STATISTICS OF THE VELOCITY GRADIENT TENSOR IN SPACE PLASMA TURBULENT FLOWS

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Consolini, Giuseppe; Marcucci, Maria Federica; Pallocchia, Giuseppe

    2015-10-10

    In the last decade, significant advances have been presented for the theoretical characterization and experimental techniques used to measure and model all of the components of the velocity gradient tensor in the framework of fluid turbulence. Here, we attempt the evaluation of the small-scale velocity gradient tensor for a case study of space plasma turbulence, observed in the Earth's magnetosheath region by the CLUSTER mission. In detail, we investigate the joint statistics P(R, Q) of the velocity gradient geometric invariants R and Q, and find that this P(R, Q) is similar to that of the low end of the inertialmore » range for fluid turbulence, with a pronounced increase in the statistics along the so-called Vieillefosse tail. In the context of hydrodynamics, this result is referred to as the dissipation/dissipation-production due to vortex stretching.« less

  5. Dispersion of acoustic surface waves by velocity gradients

    NASA Astrophysics Data System (ADS)

    Kwon, S. D.; Kim, H. C.

    1987-10-01

    The perturbation theory of Auld [Acoustic Fields and Waves in Solids (Wiley, New York, 1973), Vol. II, p. 294], which describes the effect of a subsurface gradient on the velocity dispersion of surface waves, has been modified to a simpler form by an approximation using a newly defined velocity gradient for the case of isotropic materials. The modified theory is applied to nitrogen implantation in AISI 4140 steel with a velocity gradient of Gaussian profile, and compared with dispersion data obtained by the ultrasonic right-angle technique in the frequency range from 2.4 to 14.8 MHz. The good agreement between experiments and our theory suggests that the compound layer in the subsurface region plays a dominant role in causing the dispersion of acoustic surface waves.

  6. Constraints on Lateral S Wave Velocity Gradients around the Pacific Superplume

    NASA Astrophysics Data System (ADS)

    To, A.; Romanowicz, B.

    2006-12-01

    Global shear velocity tomographic models show two large-scale low velocity structures in the lower mantle, under southern Africa and under the mid-Pacific. While tomographic models show the shape of the structures, the gradient and amplitude of the anomalies are yet to be constrained. By forward modelling of Sdiffracted phases using the Coupled Spectral ELement Method (C-SEM, Capdeville et al., 2003), we have previously shown that observed secondary phases following the Sdiff can be explained by interaction of the wavefield with sharp boundaries of the superplumes in the south Indian and south Pacific ocean (To et al., 2005). Here, we search for further constrains on velocity gradients at the border of the Pacific superplume all around the Pacific using a multi-step approach applied to a large dataset of Sdiffracted travel times and waveforms which are sensitive to the lower most mantle. We first apply our finite frequency tomographic inversion methodology (NACT, Li and Romanowicz, 1996) which provides a good starting 3D model, which in particular allows us to position the fast and slow anomalies and their boundaries quite well, as has been shown previously, but underestimates the gradients and velocity contrasts. We then perform forward modelling of Sdiff travel times, taking into account finite frequency effects, to refine the velocity contrasts and gradients and provides the next iteration 3D model. We then perform forward modelling of waveforms, down to a frequency of 0.06Hz, using C-SEM which provides final adjustments to the model. We present a model which shows that we can constrain sharp gradients on the southern and northern edges of the Pacific Superplume. To, A., B. Romanowicz, Y. Capdeville and N. Takeuchi (2005) 3D effects of sharp boundaries at the borders of the African and Pacific Superplumes: Observation and modeling. Earth and Planetary Sceince Letters, 233: 137-153 Capdeville, Y., A. To and B. Romanowicz (2003) Coupling spectral elements and

  7. Non-Axisymmetric Line Driven Disc Winds II - Full Velocity Gradient

    NASA Astrophysics Data System (ADS)

    Dyda, Sergei; Proga, Daniel

    2018-05-01

    We study non-axisymetric features of 3D line driven winds in the Sobolev approximation, where the optical depth is calculated using the full velocity gradient. We find that non-axisymmetric density features, so called clumps, form primarily at the base of the wind on super-Sobolev length scales. The density of clumps differs by a factor of ˜3 from the azimuthal average, the magnitude of their velocity dispersion is comparable to the flow velocity and they produce ˜20% variations in the column density. Clumps may be observable because differences in density produce enhancements in emission and absorption profiles or through their velocity dispersion which enhances line broadening.

  8. Statistics of velocity gradients in two-dimensional Navier-Stokes and ocean turbulence.

    PubMed

    Schorghofer, Norbert; Gille, Sarah T

    2002-02-01

    Probability density functions and conditional averages of velocity gradients derived from upper ocean observations are compared with results from forced simulations of the two-dimensional Navier-Stokes equations. Ocean data are derived from TOPEX satellite altimeter measurements. The simulations use rapid forcing on large scales, characteristic of surface winds. The probability distributions of transverse velocity derivatives from the ocean observations agree with the forced simulations, although they differ from unforced simulations reported elsewhere. The distribution and cross correlation of velocity derivatives provide clear evidence that large coherent eddies play only a minor role in generating the observed statistics.

  9. Supernova 2010ev: A reddened high velocity gradient type Ia supernova

    NASA Astrophysics Data System (ADS)

    Gutiérrez, Claudia P.; González-Gaitán, Santiago; Folatelli, Gastón; Pignata, Giuliano; Anderson, Joseph P.; Hamuy, Mario; Morrell, Nidia; Stritzinger, Maximilian; Taubenberger, Stefan; Bufano, Filomena; Olivares E., Felipe; Haislip, Joshua B.; Reichart, Daniel E.

    2016-05-01

    Aims: We present and study the spectroscopic and photometric evolution of the type Ia supernova (SN Ia) 2010ev. Methods: We obtain and analyze multiband optical light curves and optical/near-infrared spectroscopy at low and medium resolution spanning -7 days to +300 days from the B-band maximum. Results: A photometric analysis shows that SN 2010ev is a SN Ia of normal brightness with a light-curve shape of Δm15(B) = 1.12 ± 0.02 and a stretch s = 0.94 ± 0.01 suffering significant reddening. From photometric and spectroscopic analysis, we deduce a color excess of E(B - V) = 0.25 ± 0.05 and a reddening law of Rv = 1.54 ± 0.65. Spectroscopically, SN 2010ev belongs to the broad-line SN Ia group, showing stronger than average Si IIλ6355 absorption features. We also find that SN 2010ev is a high velocity gradient SN with v˙Si = 164 ± 7 km s-1 d-1. The photometric and spectral comparison with other supernovae shows that SN 2010ev has similar colors and velocities to SN 2002bo and SN 2002dj. The analysis of the nebular spectra indicates that the [Fe II]λ7155 and [Ni II]λ7378 lines are redshifted, as expected for a high velocity gradient supernova. All these common intrinsic and extrinsic properties of the high velocity gradient (HVG) group are different from the low velocity gradient (LVG) normal SN Ia population and suggest significant variety in SN Ia explosions. This paper includes data gathered with the Du Pont Telescope at Las Campanas Observatory, Chile; and the Gemini Observatory, Cerro Pachon, Chile (Gemini Program GS-2010A-Q-14). Based on observations collected at the European Organisation for Astronomical Research in the Southern Hemisphere, Chile (ESO Programme 085.D-0577).

  10. Velocity Gradients in the Intracluster Gas of the Perseus Cluster

    NASA Astrophysics Data System (ADS)

    Dupke, Renato A.; Bregman, Joel N.

    2001-02-01

    We report the results of spatially resolved X-ray spectroscopy of eight different ASCA pointings distributed symmetrically around the center of the Perseus Cluster. The outer region of the intracluster gas is roughly isothermal, with temperature ~6-7 keV and metal abundance ~0.3 solar. Spectral analysis of the central pointing is consistent with the presence of a cooling flow and a central metal abundance gradient. A significant velocity gradient is found along an axis at a position angle of ~135°, which is ~45° discrepant with the major axis of the X-ray elongation. The radial velocity difference is found to be greater than 1000 km s-1 Mpc-1 at the 90% confidence level. Simultaneous fittings of GIS 2 and 3 indicate that the velocity gradient is significant at the 95% confidence level, and the F-test rules out constant velocities at the 99% level. Intrinsic short- and long-term variations of gain are unlikely (P<0.03) to explain the velocity discrepancies.

  11. Intermittency in small-scale turbulence: a velocity gradient approach

    NASA Astrophysics Data System (ADS)

    Meneveau, Charles; Johnson, Perry

    2017-11-01

    Intermittency of small-scale motions is an ubiquitous facet of turbulent flows, and predicting this phenomenon based on reduced models derived from first principles remains an important open problem. Here, a multiple-time scale stochastic model is introduced for the Lagrangian evolution of the full velocity gradient tensor in fluid turbulence at arbitrarily high Reynolds numbers. This low-dimensional model differs fundamentally from prior shell models and other empirically-motivated models of intermittency because the nonlinear gradient self-stretching and rotation A2 term vital to the energy cascade and intermittency development is represented exactly from the Navier-Stokes equations. With only one adjustable parameter needed to determine the model's effective Reynolds number, numerical solutions of the resulting set of stochastic differential equations show that the model predicts anomalous scaling for moments of the velocity gradient components and negative derivative skewness. It also predicts signature topological features of the velocity gradient tensor such as vorticity alignment trends with the eigen-directions of the strain-rate. This research was made possible by a graduate Fellowship from the National Science Foundation and by a Grant from The Gulf of Mexico Research Initiative.

  12. Velocity gradients and reservoir volumes lessons in computational sensitivity

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Johnson, P.W.

    1995-12-31

    The sensitivity of reservoir volume estimation from depth converted geophysical time maps to the velocity gradients employed is investigated through a simple model study. The computed volumes are disconcertingly sensitive to gradients, both horizontal and vertical. The need for an accurate method of time to depth conversion is well demonstrated by the model study in which errors in velocity are magnified 40 fold in the computation of the volume. Thus if +/- 10% accuracy in the volume is desired, we must be able to estimate the velocity at the water contact with 0.25% accuracy. Put another way, if the velocitymore » is 8000 feet per second at the well then we have only +/- 20 feet per second leeway in estimating the velocity at the water contact. Very moderate horizontal and vertical gradients would typically indicate a velocity change of a few hundred feet per second if they are in the same direction. Clearly the interpreter needs to by very careful. A methodology is demonstrated which takes into account all the information that is available, velocities, tops, depositional and lithologic spatial patterns, and common sense. It is assumed that through appropriate use of check shot and other time-depth information, that the interpreter has correctly tied the reflection picks to the well tops. Such ties are ordinarily too soft for direct time-depth conversion to give adequate depth ties. The proposed method uses a common compaction law as its basis and incorporates time picks, tops and stratigraphic maps into the depth conversion process. The resulting depth map ties the known well tops in an optimum fashion.« less

  13. Multiscale analysis of the invariants of the velocity gradient tensor in isotropic turbulence

    NASA Astrophysics Data System (ADS)

    Danish, Mohammad; Meneveau, Charles

    2018-04-01

    Knowledge of local flow-topology, the patterns of streamlines around a moving fluid element as described by the velocity-gradient tensor, is useful for developing insights into turbulence processes, such as energy cascade, material element deformation, or scalar mixing. Much has been learned in the recent past about flow topology at the smallest (viscous) scales of turbulence. However, less is known at larger scales, for instance, at the inertial scales of turbulence. In this work, we present a detailed study on the scale dependence of various quantities of interest, such as the population fraction of different types of flow-topologies, the joint probability distribution of the second and third invariants of the velocity gradient tensor, and the geometrical alignment of vorticity with strain-rate eigenvectors. We perform the analysis on a simulation dataset of isotropic turbulence at Reλ=433 . While quantities appear close to scale invariant in the inertial range, we observe a "bump" in several quantities at length scales between the inertial and viscous ranges. For instance, the population fraction of unstable node-saddle-saddle flow topology shows an increase when reducing the scale from the inertial entering the viscous range. A similar bump is observed for the vorticity-strain-rate alignment. In order to document possible dynamical causes for the different trends in the viscous and inertial ranges, we examine the probability fluxes appearing in the Fokker-Plank equation governing the velocity gradient invariants. Specifically, we aim to understand whether the differences observed between the viscous and inertial range statistics are due to effects caused by pressure, subgrid-scale, or viscous stresses or various combinations of these terms. To decompose the flow into small and large scales, we mainly use a spectrally compact non-negative filter with good spatial localization properties (Eyink-Aluie filter). The analysis shows that when going from the inertial

  14. Prescribed Velocity Gradients for Highly Viscous SPH Fluids with Vorticity Diffusion.

    PubMed

    Peer, Andreas; Teschner, Matthias

    2017-12-01

    Working with prescribed velocity gradients is a promising approach to efficiently and robustly simulate highly viscous SPH fluids. Such approaches allow to explicitly and independently process shear rate, spin, and expansion rate. This can be used to, e.g., avoid interferences between pressure and viscosity solvers. Another interesting aspect is the possibility to explicitly process the vorticity, e.g., to preserve the vorticity. In this context, this paper proposes a novel variant of the prescribed-gradient idea that handles vorticity in a physically motivated way. In contrast to a less appropriate vorticity preservation that has been used in a previous approach, vorticity is diffused. The paper illustrates the utility of the vorticity diffusion. Therefore, comparisons of the proposed vorticity diffusion with vorticity preservation and additionally with vorticity damping are presented. The paper further discusses the relation between prescribed velocity gradients and prescribed velocity Laplacians which improves the intuition behind the prescribed-gradient method for highly viscous SPH fluids. Finally, the paper discusses the relation of the proposed method to a physically correct implicit viscosity formulation.

  15. Seismic Velocity Gradients Across the Transition Zone

    NASA Astrophysics Data System (ADS)

    Escalante, C.; Cammarano, F.; de Koker, N.; Piazzoni, A.; Wang, Y.; Marone, F.; Dalton, C.; Romanowicz, B.

    2006-12-01

    One-D elastic velocity models derived from mineral physics do a notoriously poor job at predicting the velocity gradients in the upper mantle transition zone, as well as some other features of models derived from seismological data. During the 2006 CIDER summer program, we computed Vs and Vp velocity profiles in the upper mantle based on three different mineral physics approaches: two approaches based on the minimization of Gibbs Free Energy (Stixrude and Lithgow-Bertelloni, 2005; Piazzoni et al., 2006) and one obtained by using experimentally determined phase diagrams (Weidner and Wang, 1998). The profiles were compared by assuming a vertical temperature profile and two end-member compositional models, the pyrolite model of Ringwood (1979) and the piclogite model of Anderson and Bass (1984). The predicted seismic profiles, which are significantly different from each other, primarily due to different choices of properties of single minerals and their extrapolation with temperature, are tested against a global dataset of P and S travel times and spheroidal and toroidal normal mode eigenfrequencies. All the models derived using a potential temperature of 1600K predict seismic velocities that are too slow in the upper mantle, suggesting the need to use a colder geotherm. The velocity gradient in the transition zone is somewhat better for piclogite than for pyrolite, possibly indicating the need to increase Ca content. The presence of stagnant slabs in the transition zone is a possible explanation for the need for 1) colder temperature and 2) increased Ca content. Future improvements in seismic profiles obtained from mineral physics will arise from better knowledge of elastic properties of upper mantle constituents and aggregates at high temperature and pressure, a better understanding of differences between thermodynamic models, and possibly the effect of water through and on Q. High resolution seismic constraints on velocity jumps at 400 and 660 km also need to be

  16. Study of the velocity gradient tensor in turbulent flow

    NASA Technical Reports Server (NTRS)

    Cheng, Wei-Ping; Cantwell, Brian

    1996-01-01

    The behavior of the velocity gradient tensor, A(ij)=delta u(i)/delta x(j), was studied using three turbulent flows obtained from direct numerical simulation The flows studies were: an inviscid calculation of the interaction between two vortex tubes, a homogeneous isotropic flow, and a temporally evolving planar wake. Self-similar behavior for each flow was obtained when A(ij) was normalized with the mean strain rate. The case of the interaction between two vortex tubes revealed a finite sized coherent structure with topological characteristics predictable by a restricted Euler model. This structure was found to evolve with the peak vorticity as the flow approached singularity. Invariants of A(ij) within this structure followed a straight line relationship of the form: gamma(sup 3)+gammaQ+R=0, where Q and R are the second and third invariants of A(ij), and the eigenvalue gamma is nearly constant over the volume of this structure. Data within this structure have local strain topology of unstable-node/saddle/saddle. The characteristics of the velocity gradient tensor and the anisotropic part of a related acceleration gradient tensor H(ij) were also studied for a homogeneous isotropic flow and a temporally evolving planar wake. It was found that the intermediate principal eigenvalue of the rate-of-strain tensor of H(ij) tended to be negative, with local strain topology of the type stable-node/saddle/saddle. There was also a preferential eigenvalue direction. The magnitude of H(ij) in the wake flow was found to be very small when data were conditioned at high local dissipation regions. This result was not observed in the relatively low Reynolds number simulation of homogeneous isotropic flow. A restricted Euler model of the evolution of A(ij) was found to reproduce many of the topological features identified in the simulations.

  17. Numerical studies of asymmetric adiabatic accretion flow - The effect of velocity gradients

    NASA Technical Reports Server (NTRS)

    Taam, Ronald E.; Fryxell, B. A.

    1989-01-01

    A numerical study of the time variation of the angular momentum and mass capture rates for a central object accreting from a uniform medium with a velocity gradient transverse to the direction of the mean flow is presented, covering a range of velocity asymmetries and Mach numbers in the incident flow. It is found that the mass accretion rate in a given evolutionary sequence varies in an irregular manner, with the matter accreting onto the central object from either a continuously moving accretion wake or from an accretion disk. The implications of the results from the study of short-term fluctuations observed in the pulse period and luminosity of X-ray pulsars are discussed.

  18. Airship stresses due to vertical velocity gradients and atmospheric turbulence

    NASA Technical Reports Server (NTRS)

    Sheldon, D.

    1975-01-01

    Munk's potential flow method is used to calculate the resultant moment experienced by an ellipsoidal airship. This method is first used to calculate the moment arising from basic maneuvers considered by early designers, and then expended to calculate the moment arising from vertical velocity gradients and atmospheric turbulence. This resultant moment must be neutralized by the transverse force of the fins. The results show that vertical velocity gradients at a height of 6000 feet in thunderstorms produce a resultant moment approximately three to four times greater than the moment produced in still air by realistic values of pitch angle or steady turning. Realistic values of atmospheric turbulence produce a moment which is significantly less than the moment produced by maneuvers in still air.

  19. Velocity Gradient Power Functional for Brownian Dynamics.

    PubMed

    de Las Heras, Daniel; Schmidt, Matthias

    2018-01-12

    We present an explicit and simple approximation for the superadiabatic excess (over ideal gas) free power functional, admitting the study of the nonequilibrium dynamics of overdamped Brownian many-body systems. The functional depends on the local velocity gradient and is systematically obtained from treating the microscopic stress distribution as a conjugate field. The resulting superadiabatic forces are beyond dynamical density functional theory and are of a viscous nature. Their high accuracy is demonstrated by comparison to simulation results.

  20. Velocity Gradient Power Functional for Brownian Dynamics

    NASA Astrophysics Data System (ADS)

    de las Heras, Daniel; Schmidt, Matthias

    2018-01-01

    We present an explicit and simple approximation for the superadiabatic excess (over ideal gas) free power functional, admitting the study of the nonequilibrium dynamics of overdamped Brownian many-body systems. The functional depends on the local velocity gradient and is systematically obtained from treating the microscopic stress distribution as a conjugate field. The resulting superadiabatic forces are beyond dynamical density functional theory and are of a viscous nature. Their high accuracy is demonstrated by comparison to simulation results.

  1. Tracing Interstellar Magnetic Field Using Velocity Gradient Technique: Application to Atomic Hydrogen Data

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Yuen, Ka Ho; Lazarian, A., E-mail: kyuen2@wisc.edu, E-mail: lazarian@astro.wisc.edu

    The advancement of our understanding of MHD turbulence opens ways to develop new techniques to probe magnetic fields. In MHD turbulence, the velocity gradients are expected to be perpendicular to magnetic fields and this fact was used by González-Casanova and Lazarian to introduce a new technique to trace magnetic fields using velocity centroid gradients (VCGs). The latter can be obtained from spectroscopic observations. We apply the technique to GALFA-H i survey data and then compare the directions of magnetic fields obtained with our technique to the direction of magnetic fields obtained using PLANCK polarization. We find an excellent correspondence betweenmore » the two ways of magnetic field tracing, which is obvious via the visual comparison and through the measuring of the statistics of magnetic field fluctuations obtained with the polarization data and our technique. This suggests that the VCGs have a potential for measuring of the foreground magnetic field fluctuations, and thus provide a new way of separating foreground and CMB polarization signals.« less

  2. Arm classification and velocity gradients in spiral galaxies

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Biviano, A.; Girardi, M.; Giuricin, G.

    1991-08-01

    On the basis of published rotation curves, velocity gradients are compiled for 94 galaxies. A significant correlation is found in this sample of galaxies between their gradients and arm classes (as given by Elmegreen and Elmegreen, 1982); galaxies with steeper curves tend to have a flocculent arm structure, and galaxies with flatter curves tend to have a grand design morphology. The correlation is true, since it is not induced by other correlations. The present result is in agreement with previous suggestions by Whitmore (1984) and with the recent result by Elmegreen and Elmegreen; it is also consistent with the predictionsmore » of density wave theory for the formation of the spiral structure. 89 refs.« less

  3. Acoustic beam control in biomimetic projector via velocity gradient

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Gao, Xiaowei; Dong, Erqian; Song, Zhongchang

    A biomimetic projector (BioP) based on computerized tomography of pygmy sperm whale's biosonar system has been designed using gradient-index (GRIN) material. The directivity of this BioP device was investigated as function of frequency and the velocity gradient of the GRIN material. A strong beam control over a broad bandwidth at the subwavelength scale has been achieved. Compared with a bare subwavelength source, the main lobe pressure of the BioP is about five times as high and the angular resolution is one order of magnitude better. Our results indicate that this BioP has excellent application potential in miniaturized underwater sonars.

  4. Acoustic beam control in biomimetic projector via velocity gradient

    NASA Astrophysics Data System (ADS)

    Gao, Xiaowei; Zhang, Yu; Cao, Wenwu; Dong, Erqian; Song, Zhongchang; Li, Songhai; Tang, Liguo; Zhang, Sai

    2016-07-01

    A biomimetic projector (BioP) based on computerized tomography of pygmy sperm whale's biosonar system has been designed using gradient-index (GRIN) material. The directivity of this BioP device was investigated as function of frequency and the velocity gradient of the GRIN material. A strong beam control over a broad bandwidth at the subwavelength scale has been achieved. Compared with a bare subwavelength source, the main lobe pressure of the BioP is about five times as high and the angular resolution is one order of magnitude better. Our results indicate that this BioP has excellent application potential in miniaturized underwater sonars.

  5. What Do the Hitomi Observations Tell Us About the Turbulent Velocities in the Perseus Cluster? Probing the Velocity Field with Mock Observations

    NASA Astrophysics Data System (ADS)

    ZuHone, J. A.; Miller, E. D.; Bulbul, E.; Zhuravleva, I.

    2018-02-01

    Hitomi made the first direct measurements of galaxy cluster gas motions in the Perseus cluster, which implied that its core is fairly “quiescent,” with velocities less than ∼200 km s‑1, despite the presence of an active galactic nucleus and sloshing cold fronts. Building on previous work, we use synthetic Hitomi/X-ray Spectrometer (SXS) observations of the hot plasma of a simulated cluster with sloshing gas motions and varying viscosity to analyze its velocity structure in a similar fashion. We find that sloshing motions can produce line shifts and widths similar to those measured by Hitomi. We find these measurements are unaffected by the value of the gas viscosity, since its effects are only manifested clearly on angular scales smaller than the SXS ∼1‧ PSF. The PSF biases the line shift of regions near the core as much as ∼40–50 km s‑1, so it is crucial to model this effect carefully. We also infer that if sloshing motions dominate the observed velocity gradient, Perseus must be observed from a line of sight that is somewhat inclined from the plane of these motions, but one that still allows the spiral pattern to be visible. Finally, we find that assuming isotropy of motions can underestimate the total velocity and kinetic energy of the core in our simulation by as much as ∼60%. However, the total kinetic energy in our simulated cluster core is still less than 10% of the thermal energy in the core, in agreement with the Hitomi observations.

  6. Dynamically balanced absolute sea level of the global ocean derived from near-surface velocity observations

    NASA Astrophysics Data System (ADS)

    Niiler, Pearn P.; Maximenko, Nikolai A.; McWilliams, James C.

    2003-11-01

    The 1992-2002 time-mean absolute sea level distribution of the global ocean is computed for the first time from observations of near-surface velocity. For this computation, we use the near-surface horizontal momentum balance. The velocity observed by drifters is used to compute the Coriolis force and the force due to acceleration of water parcels. The anomaly of horizontal pressure gradient is derived from satellite altimetry and corrects the temporal bias in drifter data distribution. NCEP reanalysis winds are used to compute the force due to Ekman currents. The mean sea level gradient force, which closes the momentum balance, is integrated for mean sea level. We find that our computation agrees, within uncertainties, with the sea level computed from the geostrophic, hydrostatic momentum balance using historical mean density, except in the Antarctic Circumpolar Current. A consistent horizontally and vertically dynamically balanced, near-surface, global pressure field has now been derived from observations.

  7. Sinking velocities of phytoplankton measured on a stable density gradient by laser scanning

    PubMed Central

    Walsby, Anthony E; Holland, Daryl P

    2005-01-01

    Two particular difficulties in measuring the sinking velocities of phytoplankton cells are preventing convection within the sedimenting medium and determining the changing depth of the cells. These problems are overcome by using a density-stabilized sedimentation column scanned by a laser. For freshwater species, a suspension of phytoplankton is layered over a vertical density gradient of Percoll solution; as the cells sink down the column their relative concentration is measured by the forward scattering of light from a laser beam that repeatedly scans up and down the column. The Percoll gradient stabilizes the column, preventing vertical mixing by convection, radiation or perturbation of density by the descending cells. Measurements were made on suspensions of 15 μm polystyrene microspheres with a density of 1050 kg m−3; the mean velocity was 6.28 μm s−1, within 1.5% of that calculated by the Stokes equation, 6.36 μm s−1. Measurements made on the filamentous cyanobacterium Planktothrix rubescens gave mean velocities within the theoretical range of values based on the range of size, shape, orientation and density of the particles in a modified Stokes equation. Measurements on marine phytoplankton may require density gradients prepared with other substances. PMID:16849271

  8. On the effect of velocity gradients on the depth of correlation in μPIV

    NASA Astrophysics Data System (ADS)

    Mustin, B.; Stoeber, B.

    2016-03-01

    The present work revisits the effect of velocity gradients on the depth of the measurement volume (depth of correlation) in microscopic particle image velocimetry (μPIV). General relations between the μPIV weighting functions and the local correlation function are derived from the original definition of the weighting functions. These relations are used to investigate under which circumstances the weighting functions are related to the curvature of the local correlation function. Furthermore, this work proposes a modified definition of the depth of correlation that leads to more realistic results than previous definitions for the case when flow gradients are taken into account. Dimensionless parameters suitable to describe the effect of velocity gradients on μPIV cross correlation are derived and visual interpretations of these parameters are proposed. We then investigate the effect of the dimensionless parameters on the weighting functions and the depth of correlation for different flow fields with spatially constant flow gradients and with spatially varying gradients. Finally this work demonstrates that the results and dimensionless parameters are not strictly bound to a certain model for particle image intensity distributions but are also meaningful when other models for particle images are used.

  9. New Models for Velocity/Pressure-Gradient Correlations in Turbulent Boundary Layers

    NASA Astrophysics Data System (ADS)

    Poroseva, Svetlana; Murman, Scott

    2014-11-01

    To improve the performance of Reynolds-Averaged Navier-Stokes (RANS) turbulence models, one has to improve the accuracy of models for three physical processes: turbulent diffusion, interaction of turbulent pressure and velocity fluctuation fields, and dissipative processes. The accuracy of modeling the turbulent diffusion depends on the order of a statistical closure chosen as a basis for a RANS model. When the Gram-Charlier series expansions for the velocity correlations are used to close the set of RANS equations, no assumption on Gaussian turbulence is invoked and no unknown model coefficients are introduced into the modeled equations. In such a way, this closure procedure reduces the modeling uncertainty of fourth-order RANS (FORANS) closures. Experimental and direct numerical simulation data confirmed the validity of using the Gram-Charlier series expansions in various flows including boundary layers. We will address modeling the velocity/pressure-gradient correlations. New linear models will be introduced for the second- and higher-order correlations applicable to two-dimensional incompressible wall-bounded flows. Results of models' validation with DNS data in a channel flow and in a zero-pressure gradient boundary layer over a flat plate will be demonstrated. A part of the material is based upon work supported by NASA under award NNX12AJ61A.

  10. Joint Analysis of GOCE Gravity Gradients Data with Seismological and Geodynamic Observations to Infer Mantle Properties

    NASA Astrophysics Data System (ADS)

    Metivier, L.; Greff-Lefftz, M.; Panet, I.; Pajot-Métivier, G.; Caron, L.

    2014-12-01

    Joint inversion of the observed geoid and seismic velocities has been commonly used to constrain the viscosity profile within the mantle as well as the lateral density variations. Recent satellite measurements of the second-order derivatives of the Earth's gravity potential give new possibilities to understand these mantle properties. We use lateral density variations in the Earth's mantle based on slab history or deduced from seismic tomography. The main uncertainties are the relationship between seismic velocity and density -the so-called density/velocity scaling factor- and the variation with depth of the density contrast between the cold slabs and the surrounding mantle, introduced here as a scaling factor with respect to a constant value. The geoid, gravity and gravity gradients at the altitude of the GOCE satellite (about 255 km) are derived using geoid kernels for given viscosity depth profiles. We assume a layered mantle model with viscosity and conversion factor constant in each layer, and we fix the viscosity of the lithosphere. We perform a Monte Carlo search for the viscosity and the density/velocity scaling factor profiles within the mantle which allow to fit the observed geoid, gravity and gradients of gravity. We test a 2-layer, a 3-layer and 4-layer mantle. For each model, we compute the posterior probability distribution of the unknown parameters, and we discuss the respective contributions of the geoid, gravity and gravity gradients in the inversion. Finally, for the best fit, we present the viscosity and scaling factor profiles obtained for the lateral density variations derived from seismic velocities and for slabs sinking into the mantle.

  11. The influence of a high pressure gradient on unsteady velocity perturbations in the case of a turbulent supersonic flow

    NASA Technical Reports Server (NTRS)

    Dussauge, J. P.; Debieve, J. F.

    1980-01-01

    The amplification or reduction of unsteady velocity perturbations under the influence of strong flow acceleration or deceleration was studied. Supersonic flows with large velocity, pressure gradients, and the conditions in which the velocity fluctuations depend on the action of the average gradients of pressure and velocity rather than turbulence, are described. Results are analyzed statistically and interpreted as a return to laminar process. It is shown that this return to laminar implies negative values in the turbulence production terms for kinetic energy. A simple geometrical representation of the Reynolds stress production is given.

  12. Microgravity Segregation in Binary Mixtures of Inelastic Spheres Driven by Velocity Fluctuation Gradients

    NASA Technical Reports Server (NTRS)

    Jenkins, James T.; Louge, Michel Y.

    1996-01-01

    We are interested in collisional granular flows of dry materials in reduced gravity. Because the particles interact through collisions, the energy of the particle velocity fluctuations plays an important role in the physics. Here we focus on the separation of grains by properties - size, for example - that is driven by spatial gradients in the fluctuation energy of the grains. The segregation of grains by size is commonly observed in geophysical flows and industrial processes. Segregation of flowing grains can also take place based on other properties, e.g. shape, mass, friction, and coefficient of restitution. Many mechanisms may be responsible for segregation; most of these are strongly influenced by gravity. Here, we outline a mechanism that is independent of gravity. This mechanism may be important but is often obscured in terrestrial grain flows. It is driven by gradients in fluctuation energy. In microgravity, the separation of grains by property will proceed slowly enough to permit flight observations to provide an unambiguous measurement of the transport coefficients associated with the segregation. In this context, we are planning a microgravity shear cell experiment that contains a mixture of two types of spherical grains. The grains will be driven to interact with two different types of boundaries on either sides of the cell. The resulting separation will be observed visually.

  13. A strongly negative shear velocity gradient and lateral variability in the lowermost mantle beneath the Pacific

    NASA Astrophysics Data System (ADS)

    Ritsema, Jeroen; Garnero, Edward; Lay, Thorne

    1997-01-01

    A new approach for constraining the seismic shear velocity structure above the core-mantle boundary is introduced, whereby SH-SKS differential travel times, amplitude ratios of SV/SKS, and Sdiff waveshapes are simultaneously modeled. This procedure is applied to the lower mantle beneath the central Pacific using da.ta from numerous deep-focus southwest Pacific earthquakes recorded in North America. We analyze 90 broadband and 248 digitized analog recordings for this source-receiver geometry. SH-SKS times are highly variable and up to 10 s larger than standard reference model predictions, indicating the presence of laterally varying low shear velocities in the study area. The travel times, however, do not constrain the depth extent or velocity gradient of the low-velocity region. SV/SKS amplitude ratios and SH waveforms are sensitive to the radial shear velocity profile, and when analyzed simultaneously with SH-SKS times, rnveal up to 3% shear velocity reductions restricted to the lowermost 190±50 km of the mantle. Our preferred model for the central-eastern Pacific region (Ml) has a strong negative gradient (with 0.5% reduction in velocity relative to the preliminary reference Earth model (PREM) at 2700 km depth and 3% reduction at 2891 km depth) and slight velocity reductions from 2000 to 2700 km depth (0-0.5% lower than PREM). Significant small-scale (100-500 km) shear velocity heterogeneity (0.5%-1%) is required to explain scatter in the differential times and amplitude ratios.

  14. Amplitude of Sdiff across Asia: effects of velocity gradient and Qs in the D'' region and the asphericity of the mantle

    NASA Astrophysics Data System (ADS)

    Kuo, Ban-Yuan

    1999-11-01

    The amplitudes of diffracted SH (S diff) normalized to SKS, together with the S diff-SKS times, were analyzed to constrain the structure of the D" region beneath Asia and the northernmost Indian Ocean. While the S diff-SKS residuals (δt; relative to the Preliminary Reference Earth model, or PREM) are consistently negative from 95° to 120°, the amplitude residuals of S diff/SKS (δ A) show two trends of distance dependence, corresponding to distinct seismic structures in two adjacent zones in D". In zone A, δ A increases significantly with distance, suggesting the presence of a negative velocity gradient in the base of the mantle. The travel time residuals independently require that the average velocity of zone A be faster than that of PREM. One-dimensional structures that reconcile both sets of constraints were sought through systematic forwarding modeling. Models with negative gradients that satisfy δt's match δ A's to an acceptable degree only if a high-quality factor ( Qs) is assumed. The preferred model for zone A has a 400-500 km thick negative gradient layer, with a ~4% velocity discontinuity at the top and Qs = 1000, an about three-fold increase from the PREM value. In zone B, the amplitude-distance curve is virtually flat, and a 200-300 km thick high-velocity layer with PREM-like gradient and Qs explains both observations well. To assess the role of mantle asphericity in δ A, we estimate the strength of focusing of the S waves into the Fresnel zone at the onset of diffraction in vertical cross-sections of 3-D tomographic models SAW12D and SKS12WM13. Both models predict stronger focusing in zone A than in zone B. The focusing effect is translated to a positive base-line shift in δ A, which, if applied to the model predictions, alleviates the need for an extremely high Qs in zone A. The simple 2-D experiment suggests that velocity gradient and the anelastic attenuation of the D" layer as well as the mantle heterogeneity all probably contribute to the

  15. Direct mapping of the temperature and velocity gradients in discs. Imaging the vertical CO snow line around IM Lupi

    NASA Astrophysics Data System (ADS)

    Pinte, C.; Ménard, F.; Duchêne, G.; Hill, T.; Dent, W. R. F.; Woitke, P.; Maret, S.; van der Plas, G.; Hales, A.; Kamp, I.; Thi, W. F.; de Gregorio-Monsalvo, I.; Rab, C.; Quanz, S. P.; Avenhaus, H.; Carmona, A.; Casassus, S.

    2018-01-01

    Accurate measurements of the physical structure of protoplanetary discs are critical inputs for planet formation models. These constraints are traditionally established via complex modelling of continuum and line observations. Instead, we present an empirical framework to locate the CO isotopologue emitting surfaces from high spectral and spatial resolution ALMA observations. We apply this framework to the disc surrounding IM Lupi, where we report the first direct, i.e. model independent, measurements of the radial and vertical gradients of temperature and velocity in a protoplanetary disc. The measured disc structure is consistent with an irradiated self-similar disc structure, where the temperature increases and the velocity decreases towards the disc surface. We also directly map the vertical CO snow line, which is located at about one gas scale height at radii between 150 and 300 au, with a CO freeze-out temperature of 21 ± 2 K. In the outer disc (>300 au), where the gas surface density transitions from a power law to an exponential taper, the velocity rotation field becomes significantly sub-Keplerian, in agreement with the expected steeper pressure gradient. The sub-Keplerian velocities should result in a very efficient inward migration of large dust grains, explaining the lack of millimetre continuum emission outside of 300 au. The sub-Keplerian motions may also be the signature of the base of an externally irradiated photo-evaporative wind. In the same outer region, the measured CO temperature above the snow line decreases to ≈15 K because of the reduced gas density, which can result in a lower CO freeze-out temperature, photo-desorption, or deviations from local thermodynamic equilibrium.

  16. Field estimates of floc dynamics and settling velocities in a tidal creek with significant along-channel gradients in velocity and SPM

    NASA Astrophysics Data System (ADS)

    Schwarz, C.; Cox, T.; van Engeland, T.; van Oevelen, D.; van Belzen, J.; van de Koppel, J.; Soetaert, K.; Bouma, T. J.; Meire, P.; Temmerman, S.

    2017-10-01

    A short-term intensive measurement campaign focused on flow, turbulence, suspended particle concentration, floc dynamics and settling velocities were carried out in a brackish intertidal creek draining into the main channel of the Scheldt estuary. We compare in situ estimates of settling velocities between a laser diffraction (LISST) and an acoustic Doppler technique (ADV) at 20 and 40 cm above bottom (cmab). The temporal variation in settling velocity estimated were compared over one tidal cycle, with a maximum flood velocity of 0.46 m s-1, a maximum horizontal ebb velocity of 0.35 m s-1 and a maximum water depth at high water slack of 2.41 m. Results suggest that flocculation processes play an important role in controlling sediment transport processes in the measured intertidal creek. During high-water slack, particles flocculated to sizes up to 190 μm, whereas at maximum flood and maximum ebb tidal stage floc sizes only reached up to 55 μm and 71 μm respectively. These large differences indicate that flocculation processes are mainly governed by turbulence-induced shear rate. In this study, we specifically recognize the importance of along-channel gradients that places constraints on the application of the acoustic Doppler technique due to conflicts with the underlying assumptions. Along-channel gradients were assessed by additional measurements at a second location and scaling arguments which could be used as an indication whether the Reynolds-flux method is applicable. We further show the potential impact of along-channel advection of flocs out of equilibrium with local hydrodynamics influencing overall floc sizes.

  17. Observation and simulation of flow on soap film induced by concentration gradient

    NASA Astrophysics Data System (ADS)

    Ohnishi, Mitsuru; Yoshihara, Shoichi; Azuma, Hisao

    The behavior of the flow and capillary wave induced on the film surface by the surfactant concentration difference is studied. Flat soap film is used as a model of thin film. The result is applicable to the case of flow by thermal gradient. The Schlieren method is used to observe the flow and the wave on the soap film. It is found that the wave velocities, in the case of a high surface tension difference, are linearly related to the square root of the surface tension difference.

  18. Formulating viscous hydrodynamics for large velocity gradients

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Pratt, Scott

    2008-02-15

    Viscous corrections to relativistic hydrodynamics, which are usually formulated for small velocity gradients, have recently been extended from Navier-Stokes formulations to a class of treatments based on Israel-Stewart equations. Israel-Stewart treatments, which treat the spatial components of the stress-energy tensor {tau}{sub ij} as dynamical objects, introduce new parameters, such as the relaxation times describing nonequilibrium behavior of the elements {tau}{sub ij}. By considering linear response theory and entropy constraints, we show how the additional parameters are related to fluctuations of {tau}{sub ij}. Furthermore, the Israel-Stewart parameters are analyzed for their ability to provide stable and physical solutions for sound waves.more » Finally, it is shown how these parameters, which are naturally described by correlation functions in real time, might be constrained by lattice calculations, which are based on path-integral formulations in imaginary time.« less

  19. Velocity Gradient Across the San Andreas Fault and Changes in Slip Behavior as Outlined by Full non Linear Tomography

    NASA Astrophysics Data System (ADS)

    Chiarabba, C.; Giacomuzzi, G.; Piana Agostinetti, N.

    2017-12-01

    The San Andreas Fault (SAF) near Parkfield is the best known fault section which exhibit a clear transition in slip behavior from stable to unstable. Intensive monitoring and decades of studies permit to identify details of these processes with a good definition of fault structure and subsurface models. Tomographic models computed so far revealed the existence of large velocity contrasts, yielding physical insight on fault rheology. In this study, we applied a recently developed full non-linear tomography method to compute Vp and Vs models which focus on the section of the fault that exhibit fault slip transition. The new tomographic code allows not to impose a vertical seismic discontinuity at the fault position, as routinely done in linearized codes. Any lateral velocity contrast found is directly dictated by the data themselves and not imposed by subjective choices. The use of the same dataset of previous tomographic studies allows a proper comparison of results. We use a total of 861 earthquakes, 72 blasts and 82 shots and the overall arrival time dataset consists of 43948 P- and 29158 S-wave arrival times, accurately selected to take care of seismic anisotropy. Computed Vp and Vp/Vs models, which by-pass the main problems related to linarized LET algorithms, excellently match independent available constraints and show crustal heterogeneities with a high resolution. The high resolution obtained in the fault surroundings permits to infer lateral changes of Vp and Vp/Vs across the fault (velocity gradient). We observe that stable and unstable sliding sections of the SAF have different velocity gradients, small and negligible in the stable slip segment, but larger than 15 % in the unstable slip segment. Our results suggest that Vp and Vp/Vs gradients across the fault control fault rheology and the attitude of fault slip behavior.

  20. Determining mineral weathering rates based on solid and solute weathering gradients and velocities: Application to biotite weathering in saprolites

    USGS Publications Warehouse

    White, A.F.

    2002-01-01

    Chemical weathering gradients are defined by the changes in the measured elemental concentrations in solids and pore waters with depth in soils and regoliths. An increase in the mineral weathering rate increases the change in these concentrations with depth while increases in the weathering velocity decrease the change. The solid-state weathering velocity is the rate at which the weathering front propagates through the regolith and the solute weathering velocity is equivalent to the rate of pore water infiltration. These relationships provide a unifying approach to calculating both solid and solute weathering rates from the respective ratios of the weathering velocities and gradients. Contemporary weathering rates based on solute residence times can be directly compared to long-term past weathering based on changes in regolith composition. Both rates incorporate identical parameters describing mineral abundance, stoichiometry, and surface area. Weathering gradients were used to calculate biotite weathering rates in saprolitic regoliths in the Piedmont of Northern Georgia, USA and in Luquillo Mountains of Puerto Rico. Solid-state weathering gradients for Mg and K at Panola produced reaction rates of 3 to 6 x 10-17 mol m-2 s-1 for biotite. Faster weathering rates of 1.8 to 3.6 ?? 10-16 mol m-2 s-1 are calculated based on Mg and K pore water gradients in the Rio Icacos regolith. The relative rates are in agreement with a warmer and wetter tropical climate in Puerto Rico. Both natural rates are three to six orders of magnitude slower than reported experimental rates of biotite weathering. ?? 2002 Elsevier Science B.V. All rights reserved.

  1. Velocity encoding with the slice select refocusing gradient for faster imaging and reduced chemical shift-induced phase errors.

    PubMed

    Middione, Matthew J; Thompson, Richard B; Ennis, Daniel B

    2014-06-01

    To investigate a novel phase-contrast MRI velocity-encoding technique for faster imaging and reduced chemical shift-induced phase errors. Velocity encoding with the slice select refocusing gradient achieves the target gradient moment by time shifting the refocusing gradient, which enables the use of the minimum in-phase echo time (TE) for faster imaging and reduced chemical shift-induced phase errors. Net forward flow was compared in 10 healthy subjects (N = 10) within the ascending aorta (aAo), main pulmonary artery (PA), and right/left pulmonary arteries (RPA/LPA) using conventional flow compensated and flow encoded (401 Hz/px and TE = 3.08 ms) and slice select refocused gradient velocity encoding (814 Hz/px and TE = 2.46 ms) at 3 T. Improved net forward flow agreement was measured across all vessels for slice select refocused gradient compared to flow compensated and flow encoded: aAo vs. PA (1.7% ± 1.9% vs. 5.8% ± 2.8%, P = 0.002), aAo vs. RPA + LPA (2.1% ± 1.7% vs. 6.0% ± 4.3%, P = 0.03), and PA vs. RPA + LPA (2.9% ± 2.1% vs. 6.1% ± 6.3%, P = 0.04), while increasing temporal resolution (35%) and signal-to-noise ratio (33%). Slice select refocused gradient phase-contrast MRI with a high receiver bandwidth and minimum in-phase TE provides more accurate and less variable flow measurements through the reduction of chemical shift-induced phase errors and a reduced TE/repetition time, which can be used to increase the temporal/spatial resolution and/or reduce breath hold durations. Copyright © 2013 Wiley Periodicals, Inc.

  2. Note on seismic hazard assessment using gradient of uplift velocities in the Turan block (Central Asia)

    NASA Astrophysics Data System (ADS)

    Jaboyedoff, M.; Derron, M.-H.; Manby, G. M.

    2005-01-01

    Uplift gradients can provide the location of highly strained zones, which can be considered to be seismic. The Turan block (Central Asia) contains zones with high gradient of uplift velocities, above the threshold 0.04mm km-1year-1. Some of these zones are associated with important seismic activity and others are not correlated with any recent important recorded earthquakes, however, recent faults scarps as well as diverted rivers may indicate a recent tectonic activity. This threshold of gradient is probably a significant rheologic property of the upper crust. On the basis of these considerations the Uzboy river area is proposed as a potential high seismic hazard zone.

  3. Monitoring of the spatio-temporal change in the interplate coupling at northeastern Japan subduction zone based on the spatial gradients of surface velocity field

    NASA Astrophysics Data System (ADS)

    Iinuma, Takeshi

    2018-04-01

    A monitoring method to grasp the spatio-temporal change in the interplate coupling in a subduction zone based on the spatial gradients of surface displacement rate fields is proposed. I estimated the spatio-temporal change in the interplate coupling along the plate boundary in northeastern (NE) Japan by applying the proposed method to the surface displacement rates based on global positioning system observations. The gradient of the surface velocities is calculated in each swath configured along the direction normal to the Japan Trench for time windows such as 0.5, 1, 2, 3 and 5 yr being shifted by one week during the period of 1997-2016. The gradient of the horizontal velocities is negative and has a large magnitude when the interplate coupling at the shallow part (less than approximately 50 km in depth) beneath the profile is strong, and the sign of the gradient of the vertical velocity is sensitive to the existence of the coupling at the deep part (greater than approximately 50 km in depth). The trench-parallel variation of the spatial gradients of a displacement rate field clearly corresponds to the trench-parallel variation of the amplitude of the interplate coupling on the plate interface, as well as the rupture areas of previous interplate earthquakes. Temporal changes in the trench-parallel variation of the spatial gradient of the displacement rate correspond to the strengthening or weakening of the interplate coupling. We can monitor the temporal change in the interplate coupling state by calculating the spatial gradients of the surface displacement rate field to some extent without performing inversion analyses with applying certain constraint conditions that sometimes cause over- and/or underestimation at areas of limited spatial resolution far from the observation network. The results of the calculation confirm known interplate events in the NE Japan subduction zone, such as the post-seismic slip of the 2003 M8.0 Tokachi-oki and 2005 M7.2 Miyagi

  4. Study of Rayleigh-Love coupling from Spatial Gradient Observation

    NASA Astrophysics Data System (ADS)

    Lin, C. J.; Hosseini, K.; Donner, S.; Vernon, F.; Wassermann, J. M.; Igel, H.

    2017-12-01

    We present a new method to study Rayleigh-Love coupling. Instead of using seismograms solely, where ground motion is recorded as function of time, we incorporate with rotation and strain, also called spatial gradient where ground is represented as function of distance. Seismic rotation and strain are intrinsic different observable wavefield so are helpful to indentify wave type and wave propagation. A Mw 7.5 earthquake on 29 March 2015 occurred in Kokopo, Papua New Guinea recorded by a dense seismic array at PFO, California are used to obtaint seismic spatial gradient. We firstly estimate time series of azimuthal direction and phase velocity of SH wave and Rayleigh wave by analyzing collocated seismograms and rotations. This result also compares with frequency wavenumber methods using a nearby ANZA seismic array. We find the direction of Rayleigh wave fits well with great-circle back azimuth during wave propagation, while the direction of Love wave deviates from that, especially when main energy of Rayleigh wave arrives. From the analysis of cross-correlation between areal strain and vertical rotation, it reveals that high coherence, either positive or negative, happens at the same time when Love wave deparate from great-circle path. We also find the observed azimuth of Love wave and polarized particle motion of Rayleigh wave fits well with the fast direction of Rayleigh wave, for the period of 50 secs. We conclude the cause of deviated azimuth of Love wave is due to Rayleigh-Love coupling, as surface wave propagates through the area with anisotropic structure.

  5. A modified Holly-Preissmann scheme for simulating sharp concentration fronts in streams with steep velocity gradients using RIV1Q

    NASA Astrophysics Data System (ADS)

    Liu, Zhao-wei; Zhu, De-jun; Chen, Yong-can; Wang, Zhi-gang

    2014-12-01

    RIV1Q is the stand-alone water quality program of CE-QUAL-RIV1, a hydraulic and water quality model developed by U.S. Army Corps of Engineers Waterways Experiment Station. It utilizes an operator-splitting algorithm and the advection term in governing equation is treated using the explicit two-point, fourth-order accurate, Holly-Preissmann scheme, in order to preserve numerical accuracy for advection of sharp gradients in concentration. In the scheme, the spatial derivative of the transport equation, where the derivative of velocity is included, is introduced to update the first derivative of dependent variable. In the stream with larger cross-sectional variation, steep velocity gradient can be easily found and should be estimated correctly. In the original version of RIV1Q, however, the derivative of velocity is approximated by a finite difference which is first-order accurate. Its leading truncation error leads to the numerical error of concentration which is related with the velocity and concentration gradients and increases with the decreasing Courant number. The simulation may also be unstable when a sharp velocity drop occurs. In the present paper, the derivative of velocity is estimated with a modified second-order accurate scheme and the corresponding numerical error of concentration decreases. Additionally, the stability of the simulation is improved. The modified scheme is verified with a hypothetical channel case and the results demonstrate that satisfactory accuracy and stability can be achieved even when the Courant number is very low. Finally, the applicability of the modified scheme is discussed.

  6. Large-scale Observations of a Subauroral Polarization Stream by Midlatitude SuperDARN Radars: Instantaneous Longitudinal Velocity Variations

    NASA Technical Reports Server (NTRS)

    Clausen, L. B. N.; Baker, J. B. H.; Sazykin, S.; Ruohoniemi, J. M.; Greenwald, R. A.; Thomas, E. J.; Shepherd, S. G.; Talaat, E. R.; Bristow, W. A.; Zheng, Y.; hide

    2012-01-01

    We present simultaneous measurements of flow velocities inside a subauroral polarization stream (SAPS) made by six midlatitude high-frequency SuperDARN radars. The instantaneous observations cover three hours of universal time and six hours of magnetic local time (MLT). From velocity variations across the field-of-view of the radars we infer the local 2D flow direction at three different longitudes. We find that the local flow direction inside the SAPS channel is remarkably constant over the course of the event. The flow speed, however, shows significant temporal and spatial variations. After correcting for the radar look direction we are able to accurately determine the dependence of the SAPS velocity on magnetic local time. We find that the SAPS velocity variation with magnetic local time is best described by an exponential function. The average velocity at 00 MLT was 1.2 km/s and it decreased with a spatial e-folding scale of two hours of MLT toward the dawn sector. We speculate that the longitudinal distribution of pressure gradients in the ring current is responsible for this dependence and find these observations in good agreement with results from ring current models. Using TEC measurements we find that the high westward velocities of the SAPS are - as expected - located in a region of low TEC values, indicating low ionospheric conductivities.

  7. In-situ Observations of Swash-zone Flow Velocities and Sediment Transport on a Steep Beach

    NASA Astrophysics Data System (ADS)

    Chardon-Maldonado, P.; Puleo, J. A.; Figlus, J.

    2014-12-01

    A 45 m scaffolding frame containing an array of instruments was installed at South Bethany Beach, Delaware, to obtain in-situ measurements in the swash zone. Six cross-shore stations were established to simultaneously measure near-bed velocity profiles, sediment concentration and water level fluctuations on a steep beach. Measurements of swash-zone hydrodynamics and morphological change were collected from February 12 to 25, 2014, following a large Nor'easter storm with surf zone significant wave height exceeding 5 m. Swash-zone flow velocities (u,v,w) were measured at each cross-shore location using a Nortek Vectrino profiling velocimeter that measured a 30 mm velocity profile at 1 mm vertical increments at 100 Hz. These velocity profiles were used to quantify the vertical flow structure over the foreshore and estimate hydrodynamic parameters such as bed shear stress and turbulent kinetic energy dissipation. Sediment concentrations were measured using optical backscatter sensors (OBS) to obtain spatio-temporal measurements during both uprush and backwash phases of the swash cycle. Cross-shore sediment transport rates at each station were estimated by taking the product of cross-shore velocity and sediment concentration. Foreshore elevations were sampled every low tide using a Leica GPS system with RTK capability. Cross-shore sediment transport rates and gradients derived from the velocities and bed shear stress estimates will be related to the observed morphological change.

  8. Propagation of the Semidiurnal Internal Tide: Phase Velocity Versus Group Velocity

    NASA Astrophysics Data System (ADS)

    Zhao, Zhongxiang

    2017-12-01

    The superposition of two waves of slightly different wavelengths has long been used to illustrate the distinction between phase velocity and group velocity. The first-mode M2 and S2 internal tides exemplify such a two-wave model in the natural ocean. The M2 and S2 tidal frequencies are 1.932 and 2 cycles per day, respectively, and their superposition forms a spring-neap cycle in the semidiurnal band. The spring-neap cycle acts like a wave, with its frequency, wave number, and phase being the differences of the M2 and S2 internal tides. The spring-neap cycle and energy of the semidiurnal internal tide propagate at the group velocity. Long-range propagation of M2 and S2 internal tides in the North Pacific is observed by satellite altimetry. Along a 3,400 km beam spanning 24°-54°N, the M2 and S2 travel times are 10.9 and 11.2 days, respectively. For comparison, it takes the spring-neap cycle 21.1 days to travel over this distance. Spatial maps of the M2 phase velocity, the S2 phase velocity, and the group velocity are determined from phase gradients of the corresponding satellite observed internal tide fields. The observed phase and group velocities agree with theoretical values estimated using the World Ocean Atlas 2013 annual-mean ocean stratification.

  9. Model of separation performance of bilinear gradients in scanning format counter-flow gradient electrofocusing techniques.

    PubMed

    Shameli, Seyed Mostafa; Glawdel, Tomasz; Ren, Carolyn L

    2015-03-01

    Counter-flow gradient electrofocusing allows the simultaneous concentration and separation of analytes by generating a gradient in the total velocity of each analyte that is the sum of its electrophoretic velocity and the bulk counter-flow velocity. In the scanning format, the bulk counter-flow velocity is varying with time so that a number of analytes with large differences in electrophoretic mobility can be sequentially focused and passed by a single detection point. Studies have shown that nonlinear (such as a bilinear) velocity gradients along the separation channel can improve both peak capacity and separation resolution simultaneously, which cannot be realized by using a single linear gradient. Developing an effective separation system based on the scanning counter-flow nonlinear gradient electrofocusing technique usually requires extensive experimental and numerical efforts, which can be reduced significantly with the help of analytical models for design optimization and guiding experimental studies. Therefore, this study focuses on developing an analytical model to evaluate the separation performance of scanning counter-flow bilinear gradient electrofocusing methods. In particular, this model allows a bilinear gradient and a scanning rate to be optimized for the desired separation performance. The results based on this model indicate that any bilinear gradient provides a higher separation resolution (up to 100%) compared to the linear case. This model is validated by numerical studies. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

  10. Statistics of pressure and pressure gradient in homogeneous isotropic turbulence

    NASA Technical Reports Server (NTRS)

    Gotoh, T.; Rogallo, R. S.

    1994-01-01

    The statistics of pressure and pressure gradient in stationary isotropic turbulence are measured within direct numerical simulations at low to moderate Reynolds numbers. It is found that the one-point pdf of the pressure is highly skewed and that the pdf of the pressure gradient is of stretched exponential form. The power spectrum of the pressure P(k) is found to be larger than the corresponding spectrum P(sub G)(k) computed from a Gaussian velocity field having the same energy spectrum as that of the DNS field. The ratio P(k)/P(sub G)(k), a measure of the pressure-field intermittence, grows with wavenumber and Reynolds number as -R(sub lambda)(exp 1/2)log(k/k(sub d)) for k less than k(sub d)/2 where k(sub d) is the Kolmogorov wavenumber. The Lagrangian correlations of pressure gradient and velocity are compared and the Lagrangian time scale of the pressure gradient is observed to be much shorter than that of the velocity.

  11. Muscle Force-Velocity Relationships Observed in Four Different Functional Tests.

    PubMed

    Zivkovic, Milena Z; Djuric, Sasa; Cuk, Ivan; Suzovic, Dejan; Jaric, Slobodan

    2017-02-01

    The aims of the present study were to investigate the shape and strength of the force-velocity relationships observed in different functional movement tests and explore the parameters depicting force, velocity and power producing capacities of the tested muscles. Twelve subjects were tested on maximum performance in vertical jumps, cycling, bench press throws, and bench pulls performed against different loads. Thereafter, both the averaged and maximum force and velocity variables recorded from individual trials were used for force-velocity relationship modeling. The observed individual force-velocity relationships were exceptionally strong (median correlation coefficients ranged from r = 0.930 to r = 0.995) and approximately linear independently of the test and variable type. Most of the relationship parameters observed from the averaged and maximum force and velocity variable types were strongly related in all tests (r = 0.789-0.991), except for those in vertical jumps (r = 0.485-0.930). However, the generalizability of the force-velocity relationship parameters depicting maximum force, velocity and power of the tested muscles across different tests was inconsistent and on average moderate. We concluded that the linear force-velocity relationship model based on either maximum or averaged force-velocity data could provide the outcomes depicting force, velocity and power generating capacity of the tested muscles, although such outcomes can only be partially generalized across different muscles.

  12. Muscle Force-Velocity Relationships Observed in Four Different Functional Tests

    PubMed Central

    Zivkovic, Milena Z.; Djuric, Sasa; Cuk, Ivan; Suzovic, Dejan; Jaric, Slobodan

    2017-01-01

    Abstract The aims of the present study were to investigate the shape and strength of the force-velocity relationships observed in different functional movement tests and explore the parameters depicting force, velocity and power producing capacities of the tested muscles. Twelve subjects were tested on maximum performance in vertical jumps, cycling, bench press throws, and bench pulls performed against different loads. Thereafter, both the averaged and maximum force and velocity variables recorded from individual trials were used for force–velocity relationship modeling. The observed individual force-velocity relationships were exceptionally strong (median correlation coefficients ranged from r = 0.930 to r = 0.995) and approximately linear independently of the test and variable type. Most of the relationship parameters observed from the averaged and maximum force and velocity variable types were strongly related in all tests (r = 0.789-0.991), except for those in vertical jumps (r = 0.485-0.930). However, the generalizability of the force-velocity relationship parameters depicting maximum force, velocity and power of the tested muscles across different tests was inconsistent and on average moderate. We concluded that the linear force-velocity relationship model based on either maximum or averaged force-velocity data could provide the outcomes depicting force, velocity and power generating capacity of the tested muscles, although such outcomes can only be partially generalized across different muscles. PMID:28469742

  13. How to Reconcile the Observed Velocity Function of Galaxies with Theory

    NASA Astrophysics Data System (ADS)

    Brooks, Alyson M.; Papastergis, Emmanouil; Christensen, Charlotte R.; Governato, Fabio; Stilp, Adrienne; Quinn, Thomas R.; Wadsley, James

    2017-11-01

    Within a Λ cold dark matter (ΛCDM) scenario, we use high-resolution cosmological simulations spanning over four orders of magnitude in galaxy mass to understand the deficit of dwarf galaxies in observed velocity functions (VFs). We measure velocities in as similar a way as possible to observations, including generating mock H I data cubes for our simulated galaxies. We demonstrate that this apples-to-apples comparison yields an “observed” VF in agreement with observations, reconciling the large number of low-mass halos expected in a ΛCDM cosmological model with the low number of observed dwarfs at a given velocity. We then explore the source of the discrepancy between observations and theory and conclude that the dearth of observed dwarf galaxies is primarily explained by two effects. The first effect is that galactic rotational velocities derived from the H I linewidth severely underestimate the maximum halo velocity. The second effect is that a large fraction of halos at the lowest masses are too faint to be detected by current galaxy surveys. We find that cored DM density profiles can contribute to the lower observed velocity of galaxies but only for galaxies in which the velocity is measured interior to the size of the core (˜3 kpc).

  14. ASCA observation of NGC 4636: Dark matter and metallicity gradient

    NASA Technical Reports Server (NTRS)

    Mushotzky, R. F.; Loewenstein, M.; Awaki, H.; Makishima, K.; Matsushita, K.; Matsumoto, H.

    1994-01-01

    We present our analysis of ASCA PV phase observation of the elliptical galaxy NGC 4636. Solid state imaging spectrometer (SIS) spectra in six concentric annuli centered on NGC 4636 are used to derive temperature, metallicity, and column density profiles for the hot interstellar medium. Outside of the central 3 min the temperature is roughly constant at approximately 0.85 keV, while the metallicity decreases from greater than 0.36 solar at the center to less than 0.12 solar at R approximately 9 min. The implications of this gradient for elliptical galaxy formation and the enrichment of intracluster gas are discussed. We derive a detailed mass profile consistent with the stellar velocity dispersion and with ROSAT position sensitive proportional counter (PSPC) and ASCA SIS X-ray temperature profiles. We find that NGC 4636 becomes dark matter dominated at roughly the de Vaucouleurs radius, and, at r approximately 100 kpc, the ratio of dark to luminous matter density is approximately 80 and solar mass/solar luminosity approximately equal to 150. Evidence for the presence of a cooling flow is also discussed.

  15. P wave velocity of Proterozoic upper mantle beneath central and southern Asia

    NASA Astrophysics Data System (ADS)

    Nyblade, Andrew A.; Vogfjord, Kristin S.; Langston, Charles A.

    1996-05-01

    P wave velocity structure of Proterozoic upper mantle beneath central and southern Africa was investigated by forward modeling of Pnl waveforms from four moderate size earthquakes. The source-receiver path of one event crosses central Africa and lies outside the African superswell while the source-receiver paths for the other events cross Proterozoic lithosphere within southern Africa, inside the African superswell. Three observables (Pn waveshape, PL-Pn time, and Pn/PL amplitude ratio) from the Pnl waveform were used to constrain upper mantle velocity models in a grid search procedure. For central Africa, synthetic seismograms were computed for 5880 upper mantle models using the generalized ray method and wavenumber integration; synthetic seismograms for 216 models were computed for southern Africa. Successful models were taken as those whose synthetic seismograms had similar waveshapes to the observed waveforms, as well as PL-Pn times within 3 s of the observed times and Pn/PL amplitude ratios within 30% of the observed ratio. Successful models for central Africa yield a range of uppermost mantle velocity between 7.9 and 8.3 km s-1, velocities between 8.3 and 8.5 km s-1 at a depth of 200 km, and velocity gradients that are constant or slightly positive. For southern Africa, successful models yield uppermost mantle velocities between 8.1 and 8.3 km s-1, velocities between 7.9 and 8.4 km s-1 at a depth of 130 km, and velocity gradients between -0.001 and 0.001 s-1. Because velocity gradients are controlled strongly by structure at the bottoming depths for Pn waves, it is not easy to compare the velocity gradients obtained for central and southern Africa. For central Africa, Pn waves turn at depths of about 150-200 km, whereas for southern Africa they bottom at ˜100-150 km depth. With regard to the origin of the African superswell, our results do not have sufficient resolution to test hypotheses that invoke simple lithospheric reheating. However, our models are not

  16. Pickup Ion Velocity Distributions at Titan: Effects of Spatial Gradients

    NASA Technical Reports Server (NTRS)

    Hartle, R. E.; Sittler, E. C.

    2004-01-01

    The principle source of pickup ions at Titan is its neutral exosphere, extending well above the ionopause into the magnetosphere of Saturn or the solar wind, depending on the moon's orbital position. Thermal and nonthermal processes in the thermosphere generate the distribution of neutral atoms and molecules in the exosphere. The combination of these processes and the range of mass numbers, 1 to over 28, contribute to an exospheric source structure that produces pickup ions with gyroradii that are much larger or smaller than the corresponding scale heights of their neutral sources. The resulting phase space distributions are dependent on the spatial structure of the exosphere as well as that of the magnetic field and background plasma. When the pickup ion gyroradius is less than the source gas scale height, the pickup ion velocity distribution is characterized by a sharp cutoff near the maximum speed, which is twice that of the ambient plasma times the sine of the angle between the magnetic field and the flow velocity. This was the case for pickup H(sup +) ions identified during the Voyager 1 flyby. In contrast, as the gyroradius becomes much larger than the scale height, the peak of the velocity distribution in the source region recedes from the maximum speed. Iri addition, the amplitude of the distribution near the maximum speed decreases. These more beam like distributions of heavy ions were not observed from Voyager 1 , but should be observable by more sensitive instruments on future spacecraft, including Cassini. The finite gyroradius effects in the pickup ion velocity distributions are studied by including in the analysis the possible range of spatial structures in the neutral exosphere and background plasma.

  17. Experimental Study on the Flow Regimes and Pressure Gradients of Air-Oil-Water Three-Phase Flow in Horizontal Pipes

    PubMed Central

    Al-Hadhrami, Luai M.; Shaahid, S. M.; Tunde, Lukman O.; Al-Sarkhi, A.

    2014-01-01

    An experimental investigation has been carried out to study the flow regimes and pressure gradients of air-oil-water three-phase flows in 2.25 ID horizontal pipe at different flow conditions. The effects of water cuts, liquid and gas velocities on flow patterns and pressure gradients have been studied. The experiments have been conducted at 20°C using low viscosity Safrasol D80 oil, tap water and air. Superficial water and oil velocities were varied from 0.3 m/s to 3 m/s and air velocity varied from 0.29 m/s to 52.5 m/s to cover wide range of flow patterns. The experiments were performed for 10% to 90% water cuts. The flow patterns were observed and recorded using high speed video camera while the pressure drops were measured using pressure transducers and U-tube manometers. The flow patterns show strong dependence on water fraction, gas velocities, and liquid velocities. The observed flow patterns are stratified (smooth and wavy), elongated bubble, slug, dispersed bubble, and annular flow patterns. The pressure gradients have been found to increase with the increase in gas flow rates. Also, for a given superficial gas velocity, the pressure gradients increased with the increase in the superficial liquid velocity. The pressure gradient first increases and then decreases with increasing water cut. In general, phase inversion was observed with increase in the water cut. The experimental results have been compared with the existing unified Model and a good agreement has been noticed. PMID:24523645

  18. Experimental study on the flow regimes and pressure gradients of air-oil-water three-phase flow in horizontal pipes.

    PubMed

    Al-Hadhrami, Luai M; Shaahid, S M; Tunde, Lukman O; Al-Sarkhi, A

    2014-01-01

    An experimental investigation has been carried out to study the flow regimes and pressure gradients of air-oil-water three-phase flows in 2.25 ID horizontal pipe at different flow conditions. The effects of water cuts, liquid and gas velocities on flow patterns and pressure gradients have been studied. The experiments have been conducted at 20 °C using low viscosity Safrasol D80 oil, tap water and air. Superficial water and oil velocities were varied from 0.3 m/s to 3 m/s and air velocity varied from 0.29 m/s to 52.5 m/s to cover wide range of flow patterns. The experiments were performed for 10% to 90% water cuts. The flow patterns were observed and recorded using high speed video camera while the pressure drops were measured using pressure transducers and U-tube manometers. The flow patterns show strong dependence on water fraction, gas velocities, and liquid velocities. The observed flow patterns are stratified (smooth and wavy), elongated bubble, slug, dispersed bubble, and annular flow patterns. The pressure gradients have been found to increase with the increase in gas flow rates. Also, for a given superficial gas velocity, the pressure gradients increased with the increase in the superficial liquid velocity. The pressure gradient first increases and then decreases with increasing water cut. In general, phase inversion was observed with increase in the water cut. The experimental results have been compared with the existing unified Model and a good agreement has been noticed.

  19. Influences of a temperature gradient and fluid inertia on acoustic streaming in a standing wave.

    PubMed

    Thompson, Michael W; Atchley, Anthony A; Maccarone, Michael J

    2005-04-01

    Following the experimental method of Thompson and Atchley [J. Acoust. Soc. Am. 117, 1828-1838 (2005)] laser Doppler anemometry (LDA) is used to investigate the influences of a thermoacoustically induced axial temperature gradient and of fluid inertia on the acoustic streaming generated in a cylindrical standing-wave resonator filled with air driven sinusoidally at a frequency of 308 Hz. The axial component of Lagrangian streaming velocity is measured along the resonator axis and across the diameter at acoustic-velocity amplitudes of 2.7, 4.3, 6.1, and 8.6 m/s at the velocity antinodes. The magnitude of the axial temperature gradient along the resonator wall is varied between approximately 0 and 8 K/m by repeating measurements with the resonator either surrounded by a water jacket, suspended within an air-filled tank, or wrapped in foam insulation. A significant correlation is observed between the temperature gradient and the behavior of the streaming: as the magnitude of the temperature gradient increases, the magnitude of the streaming decreases and the shape of the streaming cell becomes increasingly distorted. The observed steady-state streaming velocities are not in agreement with any available theory.

  20. Lattice Boltzmann study on Kelvin-Helmholtz instability: roles of velocity and density gradients.

    PubMed

    Gan, Yanbiao; Xu, Aiguo; Zhang, Guangcai; Li, Yingjun

    2011-05-01

    A two-dimensional lattice Boltzmann model with 19 discrete velocities for compressible fluids is proposed. The fifth-order weighted essentially nonoscillatory (5th-WENO) finite difference scheme is employed to calculate the convection term of the lattice Boltzmann equation. The validity of the model is verified by comparing simulation results of the Sod shock tube with its corresponding analytical solutions [G. A. Sod, J. Comput. Phys. 27, 1 (1978).]. The velocity and density gradient effects on the Kelvin-Helmholtz instability (KHI) are investigated using the proposed model. Sharp density contours are obtained in our simulations. It is found that the linear growth rate γ for the KHI decreases by increasing the width of velocity transition layer D(v) but increases by increasing the width of density transition layer D(ρ). After the initial transient period and before the vortex has been well formed, the linear growth rates γ(v) and γ(ρ), vary with D(v) and D(ρ) approximately in the following way, lnγ(v)=a-bD(v) and γ(ρ)=c+elnD(ρ)(D(ρ)D(ρ)(E) the linear growth rate γ(ρ) does not vary significantly any more. One can use the hybrid effects of velocity and density transition layers to stabilize the KHI. Our numerical simulation results are in general agreement with the analytical results [L. F. Wang et al., Phys. Plasma 17, 042103 (2010)]. © 2011 American Physical Society

  1. Compressional velocities from multichannel refraction arrivals on Georges Bank: northwest Atlantic Ocean

    USGS Publications Warehouse

    McGinnis, L. D.; Otis, R. M.

    1979-01-01

    Velocities were obtained from unreversed, refracted arrivals on analog records from a 48‐channel, 3.6-km hydrophone cable (3.89 km from the airgun array to the last hydrophone array). Approximately 200 records were analyzed along 1500 km of ship track on Georges Bank, northwest Atlantic Ocean, to obtain regional sediment velocity distribution to a depth of 1.4 km below sea level. This technique provides nearly continuous coverage of refraction velocities and vertical velocity gradients. Because of the length of the hydrophone cable and the vertical velocity gradients, the technique is applicable only to the Continental Shelf and the shallower parts of the Continental Slope in water depths less than 300 m. Sediment diagenesis, the influence of overburden pressure on compaction, lithology, density, and porosity are inferred from these data. Velocities of the sediment near the water‐sediment interface range from less than 1500 m/sec on the north edge of Georges Bank to 1830 m/sec for glacial deposits in the northcentral part of the bank. Velocity gradients in the upper 400 m range from 1.0km/sec/km(sec−1) on the south edge of the bank to 1.7sec−1 on the north. Minimum gradients of 0.8sec−1 were observed south of Nantucket Island. Velocities and velocity gradients are explained in relation to physical properties of the Cretaceous, Tertiary, and Pleistocene sediments. Isovelocity contours at 100-m/sec intervals are nearly horizontal in the upper 400 m. Isovelocity contours at greater depths show a greater difference from a mean depth because of the greater structural and lithological variation. Bottom densities inferred from the velocities range from 1.7 to 1.9g/cm3 and porosities range from 48 to 62 percent. The most significant factor controlling velocity distribution on Georges Bank is overburden pressure and resulting compaction. From the velocity data we conclude that Georges Bank has been partially overridden by a continental ice sheet.

  2. Synthetic velocity gradient tensors and the identification of statistically significant aspects of the structure of turbulence

    NASA Astrophysics Data System (ADS)

    Keylock, Christopher J.

    2017-08-01

    A method is presented for deriving random velocity gradient tensors given a source tensor. These synthetic tensors are constrained to lie within mathematical bounds of the non-normality of the source tensor, but we do not impose direct constraints upon scalar quantities typically derived from the velocity gradient tensor and studied in fluid mechanics. Hence, it becomes possible to ask hypotheses of data at a point regarding the statistical significance of these scalar quantities. Having presented our method and the associated mathematical concepts, we apply it to homogeneous, isotropic turbulence to test the utility of the approach for a case where the behavior of the tensor is understood well. We show that, as well as the concentration of data along the Vieillefosse tail, actual turbulence is also preferentially located in the quadrant where there is both excess enstrophy (Q>0 ) and excess enstrophy production (R<0 ). We also examine the topology implied by the strain eigenvalues and find that for the statistically significant results there is a particularly strong relative preference for the formation of disklike structures in the (Q<0 ,R<0 ) quadrant. With the method shown to be useful for a turbulence that is already understood well, it should be of even greater utility for studying complex flows seen in industry and the environment.

  3. Seismic velocity and attenuation structures in the Earth's inner core

    NASA Astrophysics Data System (ADS)

    Yu, Wen-Che

    2007-12-01

    I study seismic velocity and attenuation structures in the top 400 km of the Earth's inner core along equatorial paths, velocity-attenuation relationship, and seismic anisotropy in the top of the inner core beneath Africa. Seismic observations exhibit "east-west" hemispheric differences in seismic velocity, attenuation, and anisotropy. Joint modeling of the PKiKP-PKIKP and PKPbc-PKIKP phases is used to constrain seismic velocity and attenuation structures in the top 400 km of the inner core for the eastern and western hemispheres. The velocity and attenuation models for the western hemisphere are simple, having a constant velocity gradient and a Q value of 600 in the top 400 km of the inner core. The velocity and attenuation models for the eastern hemisphere appear complex. The velocity model for the eastern hemisphere has a small velocity gradient in the top 235 km, a steeper velocity gradient at the depth range of 235 - 375 km, and a gradient similar to PREM in the deeper portion of the inner core. The attenuation model for the eastern hemisphere has a Q value of 300 in the top 300 km and a Q value of 600 in the deeper portion of the inner core. The study of velocity-attenuation relationship reveals that inner core is anisotropic in both velocity and attenuation, and the direction of high attenuation corresponding to that of high velocity. I hypothesize that the hexagonal close packed (hcp) iron crystal is anisotropic in attenuation, with the axis of high attenuation corresponding to that of high velocity. Anisotropy in the top of the inner core beneath Africa is complex. Beneath eastern Africa, the thickness of the isotropic upper inner core is about 0 km. Beneath central and western Africa, the thickness of the isotropic upper inner core increases from 20 to 50 km. The velocity increase across the isotropic upper inner core and anisotropic lower inner core boundary is sharp, laterally varying from 1.6% - 2.2%. The attenuation model has a Q value of 600 for the

  4. First seismic shear wave velocity profile of the lunar crust as extracted from the Apollo 17 active seismic data by wavefield gradient analysis

    NASA Astrophysics Data System (ADS)

    Sollberger, David; Schmelzbach, Cedric; Robertsson, Johan O. A.; Greenhalgh, Stewart A.; Nakamura, Yosio; Khan, Amir

    2016-04-01

    We present a new seismic velocity model of the shallow lunar crust, including, for the first time, shear wave velocity information. So far, the shear wave velocity structure of the lunar near-surface was effectively unconstrained due to the complexity of lunar seismograms. Intense scattering and low attenuation in the lunar crust lead to characteristic long-duration reverberations on the seismograms. The reverberations obscure later arriving shear waves and mode conversions, rendering them impossible to identify and analyze. Additionally, only vertical component data were recorded during the Apollo active seismic experiments, which further compromises the identification of shear waves. We applied a novel processing and analysis technique to the data of the Apollo 17 lunar seismic profiling experiment (LSPE), which involved recording seismic energy generated by several explosive packages on a small areal array of four vertical component geophones. Our approach is based on the analysis of the spatial gradients of the seismic wavefield and yields key parameters such as apparent phase velocity and rotational ground motion as a function of time (depth), which cannot be obtained through conventional seismic data analysis. These new observables significantly enhance the data for interpretation of the recorded seismic wavefield and allow, for example, for the identification of S wave arrivals based on their lower apparent phase velocities and distinct higher amount of generated rotational motion relative to compressional (P-) waves. Using our methodology, we successfully identified pure-mode and mode-converted refracted shear wave arrivals in the complex LSPE data and derived a P- and S-wave velocity model of the shallow lunar crust at the Apollo 17 landing site. The extracted elastic-parameter model supports the current understanding of the lunar near-surface structure, suggesting a thin layer of low-velocity lunar regolith overlying a heavily fractured crust of basaltic

  5. The effects of thermal gradients on the Mars Observer Camera primary mirror

    NASA Technical Reports Server (NTRS)

    Applewhite, Roger W.; Telkamp, Arthur R.

    1992-01-01

    The paper discusses the effect of thermal gradients on the optical performance of the primary mirror of Mars Observer Camera (MOC), which will be launched on the Mars Observer spacecraft in September 1992. It was found that mild temperature gradients can have a large effect on the mirror surface figure, even for relatively low coefficient-of-thermal-expansion materials. However, in the case of the MOC primary mirror, it was found that the radius of curvature (ROC) of the reflective surface of the mirror changed in a nearly linear fashion with the radial temperature gradient, with little additional aberration. A solid-state ROC controller using the thermal gradient effect was implemented and verified.

  6. Concentration and Velocity Gradients in Fluidized Beds

    NASA Technical Reports Server (NTRS)

    McClymer, James P.

    2003-01-01

    In this work we focus on the height dependence of particle concentration, average velocity components, fluctuations in these velocities and, with the flow turned off, the sedimentation velocity. The latter quantities are measured using Particle Imaging Velocimetry (PIV). The PIV technique uses a 1-megapixel camera to capture two time-displaced images of particles in the bed. The depth of field of the imaging system is approximately 0.5 cm. The camera images a region with characteristic length of 2.6 cm for the small particles and 4.7 cm. for the large particles. The local direction of particle flow is determined by calculating the correlation function for sub-regions of 32 x 32 pixels. The velocity vector map is created from this correlation function using the time between images (we use 15 to 30 ms). The software is sensitive variations of 1/64th of a pixel. We produce velocity maps at various heights, each consisting of 3844 velocities. We break this map into three vertical zones for increased height information. The concentration profile is measured using an expanded (1 cm diameter) linearly polarized HeNe Laser incident on the fluidized bed. A COHU camera (gamma=1, AGC off) with a lens and a polarizer images the transmitted linearly polarized light to minimize the effects of multiply scattered light. The intensity profile (640 X 480 pixels) is well described by a Gaussian fit and the height of the Gaussian is used to characterize the concentration. This value is compared to the heights found for known concentrations. The sedimentation velocity is estimated using by imaging a region near the bottom of the bed and using PIV to measure the velocity as a function of time. With a nearly uniform concentration profile, the time can be converted to height information. The stable fluidized beds are made from large pseudo-monodisperse particles (silica spheres with radii (250-300) microns and (425-500) microns) dispersed in a glycerin/water mix. The Peclet number is

  7. Experiments on the Motion of Drops on a Horizontal Solid Surface due to a Wettability Gradient

    NASA Technical Reports Server (NTRS)

    Moumen, Nadjoua; Subramanian, R, Shankar; MLaughlin, john B.

    2006-01-01

    Results from experiments performed on the motion of drops of tetraethylene glycol in a wettability gradient present on a silicon surface are reported and compared with predictions from a recently developed theoretical model. The gradient in wettability was formed by exposing strips cut from a silicon wafer to decyltrichlorosiland vapors. Video images of the drops captured during the experiments were subsequently analyzed for drop size and velocity as functions of position along the gradient. In separate experiments on the same strips, the static contact angle formed by small drops was measured and used to obtain the local wettability gradient to which a drop is subjected. The velocity of the drops was found to be a strong function of position along the gradient. A quasi-steady theoretical model that balances the local hydrodynamic resistance with the local driving force generally describes the observations; possible reasons for the remaining discrepancies are discussed. It is shown that a model in which the driving force is reduced to accomodate the hysteresis effect inferred from the data is able to remove most of the discrepancy between the observed and predicted velocities.

  8. Observations of velocity shear driven plasma turbulence

    NASA Technical Reports Server (NTRS)

    Kintner, P. M., Jr.

    1976-01-01

    Electrostatic and magnetic turbulence observations from HAWKEYE-1 during the low altitude portion of its elliptical orbit over the Southern Hemisphere are presented. The magnetic turbulence is confined near the auroral zone and is similar to that seen at higher altitudes by HEOS-2 in the polar cusp. The electrostatic turbulence is composed of a background component with a power spectral index of 1.89 + or - .26 and an intense component with a power spectral index of 2.80 + or - .34. The intense electrostatic turbulence and the magnetic turbulence correlate with velocity shears in the convective plasma flow. Since velocity shear instabilities are most unstable to wave vectors perpendicular to the magnetic field, the shear correlated turbulence is anticipated to be two dimensional in character and to have a power spectral index of 3 which agrees with that observed in the intense electrostatic turbulence.

  9. Combinational concentration gradient confinement through stagnation flow.

    PubMed

    Alicia, Toh G G; Yang, Chun; Wang, Zhiping; Nguyen, Nam-Trung

    2016-01-21

    Concentration gradient generation in microfluidics is typically constrained by two conflicting mass transport requirements: short characteristic times (τ) for precise temporal control of concentration gradients but at the expense of high flow rates and hence, high flow shear stresses (σ). To decouple the limitations from these parameters, here we propose the use of stagnation flows to confine concentration gradients within large velocity gradients that surround the stagnation point. We developed a modified cross-slot (MCS) device capable of feeding binary and combinational concentration sources in stagnation flows. We show that across the velocity well, source-sink pairs can form permanent concentration gradients. As source-sink concentration pairs are continuously supplied to the MCS, a permanently stable concentration gradient can be generated. Tuning the flow rates directly controls the velocity gradients, and hence the stagnation point location, allowing the confined concentration gradient to be focused. In addition, the flow rate ratio within the MCS rapidly controls (τ ∼ 50 ms) the location of the stagnation point and the confined combinational concentration gradients at low flow shear (0.2 Pa < σ < 2.9 Pa). The MCS device described in this study establishes the method for using stagnation flows to rapidly generate and position low shear combinational concentration gradients for shear sensitive biological assays.

  10. Observations of the scale-dependent turbulence and evaluation of the flux-gradient relationship for sensible heat for a closed Douglas-Fir canopy in very weak wind conditions

    DOE PAGES

    Vickers, D.; Thomas, C.

    2014-05-13

    Observations of the scale-dependent turbulent fluxes and variances above, within and beneath a tall closed Douglas-Fir canopy in very weak winds are examined. The daytime subcanopy vertical velocity spectra exhibit a double-peak structure with peaks at time scales of 0.8 s and 51.2 s. A double-peak structure is also observed in the daytime subcanopy heat flux cospectra. The daytime momentum flux cospectra inside the canopy and in the subcanopy are characterized by a relatively large cross-wind component, likely due to the extremely light and variable winds, such that the definition of a mean wind direction, and subsequent partitioning of themore » momentum flux into along- and cross-wind components, has little physical meaning. Positive values of both momentum flux components in the subcanopy contribute to upward transfer of momentum, consistent with the observed mean wind speed profile. In the canopy at night at the smallest resolved scales, we find relatively large momentum fluxes (compared to at larger scales), and increasing vertical velocity variance with decreasing time scale, consistent with very small eddies likely generated by wake shedding from the canopy elements that transport momentum but not heat. We find unusually large values of the velocity aspect ratio within the canopy, consistent with enhanced suppression of the horizontal wind components compared to the vertical by the canopy. The flux-gradient approach for sensible heat flux is found to be valid for the subcanopy and above-canopy layers when considered separately; however, single source approaches that ignore the canopy fail because they make the heat flux appear to be counter-gradient when in fact it is aligned with the local temperature gradient in both the subcanopy and above-canopy layers. Modeled sensible heat fluxes above dark warm closed canopies are likely underestimated using typical values of the Stanton number.« less

  11. Observations of the scale-dependent turbulence and evaluation of the flux-gradient relationship for sensible heat for a closed Douglas-Fir canopy in very weak wind conditions

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Vickers, D.; Thomas, C.

    Observations of the scale-dependent turbulent fluxes and variances above, within and beneath a tall closed Douglas-Fir canopy in very weak winds are examined. The daytime subcanopy vertical velocity spectra exhibit a double-peak structure with peaks at time scales of 0.8 s and 51.2 s. A double-peak structure is also observed in the daytime subcanopy heat flux cospectra. The daytime momentum flux cospectra inside the canopy and in the subcanopy are characterized by a relatively large cross-wind component, likely due to the extremely light and variable winds, such that the definition of a mean wind direction, and subsequent partitioning of themore » momentum flux into along- and cross-wind components, has little physical meaning. Positive values of both momentum flux components in the subcanopy contribute to upward transfer of momentum, consistent with the observed mean wind speed profile. In the canopy at night at the smallest resolved scales, we find relatively large momentum fluxes (compared to at larger scales), and increasing vertical velocity variance with decreasing time scale, consistent with very small eddies likely generated by wake shedding from the canopy elements that transport momentum but not heat. We find unusually large values of the velocity aspect ratio within the canopy, consistent with enhanced suppression of the horizontal wind components compared to the vertical by the canopy. The flux-gradient approach for sensible heat flux is found to be valid for the subcanopy and above-canopy layers when considered separately; however, single source approaches that ignore the canopy fail because they make the heat flux appear to be counter-gradient when in fact it is aligned with the local temperature gradient in both the subcanopy and above-canopy layers. Modeled sensible heat fluxes above dark warm closed canopies are likely underestimated using typical values of the Stanton number.« less

  12. Connections between density, wall-normal velocity, and coherent structure in a heated turbulent boundary layer

    NASA Astrophysics Data System (ADS)

    Saxton-Fox, Theresa; Gordeyev, Stanislav; Smith, Adam; McKeon, Beverley

    2015-11-01

    Strong density gradients associated with turbulent structure were measured in a mildly heated turbulent boundary layer using an optical sensor (Malley probe). The Malley probe measured index of refraction gradients integrated along the wall-normal direction, which, due to the proportionality of index of refraction and density in air, was equivalently an integral measure of density gradients. The integral output was observed to be dominated by strong, localized density gradients. Conditional averaging and Pearson correlations identified connections between the streamwise gradient of density and the streamwise gradient of wall-normal velocity. The trends were suggestive of a process of pick-up and transport of heat away from the wall. Additionally, by considering the density field as a passive marker of structure, the role of the wall-normal velocity in shaping turbulent structure in a sheared flow was examined. Connections were developed between sharp gradients in the density and flow fields and strong vertical velocity fluctuations. This research is made possible by the Department of Defense through the National Defense & Engineering Graduate Fellowship (NDSEG) Program and by the Air Force Office of Scientific Research Grant # FA9550-12-1-0060.

  13. Geodetic observations of ice flow velocities over the southern part of subglacial Lake Vostok, Antarctica, and their glaciological implications

    NASA Astrophysics Data System (ADS)

    Wendt, Jens; Dietrich, Reinhard; Fritsche, Mathias; Wendt, Anja; Yuskevich, Alexander; Kokhanov, Andrey; Senatorov, Anton; Lukin, Valery; Shibuya, Kazuo; Doi, Koichiro

    2006-09-01

    In the austral summer seasons 2001/02 and 2002/03, Global Positioning System (GPS) data were collected in the vicinity of Vostok Station to determine ice flow velocities over Lake Vostok. Ten GPS sites are located within a radius of 30km around Vostok Station on floating ice as well as on grounded ice to the east and to the west of the lake. Additionally, a local deformation network around the ice core drilling site 5G-1 was installed. The derived ice flow velocity for Vostok Station is 2.00ma-1 +/- 0.01ma-1. Along the flowline of Vostok Station an extension rate of about 10-5a-1 (equivalent to 1cm km-1 a-1) was determined. This significant velocity gradient results in a new estimate of 28700 years for the transit time of an ice particle along the Vostok flowline from the bedrock ridge in the southwest of the lake to the eastern shoreline. With these lower velocities compared to earlier studies and, hence, larger transit times the basal accretion rate is estimated to be 4mma-1 along a portion of the Vostok flowline. An assessment of the local accretion rate at Vostok Station using the observed geodetic quantities yields an accretion rate in the same order of magnitude. Furthermore, the comparison of our geodetic observations with results inferred from ice-penetrating radar data indicates that the ice flow may not have changed significantly for several thousand years.

  14. The Role of the Velocity Gradient in Laminar Convective Heat Transfer through a Tube with a Uniform Wall Heat Flux

    ERIC Educational Resources Information Center

    Wang, Liang-Bi; Zhang, Qiang; Li, Xiao-Xia

    2009-01-01

    This paper aims to contribute to a better understanding of convective heat transfer. For this purpose, the reason why thermal diffusivity should be placed before the Laplacian operator of the heat flux, and the role of the velocity gradient in convective heat transfer are analysed. The background to these analyses is that, when the energy…

  15. Magnetospheric Multiscale Observation of Plasma Velocity-Space Cascade: Hermite Representation and Theory.

    PubMed

    Servidio, S; Chasapis, A; Matthaeus, W H; Perrone, D; Valentini, F; Parashar, T N; Veltri, P; Gershman, D; Russell, C T; Giles, B; Fuselier, S A; Phan, T D; Burch, J

    2017-11-17

    Plasma turbulence is investigated using unprecedented high-resolution ion velocity distribution measurements by the Magnetospheric Multiscale mission (MMS) in the Earth's magnetosheath. This novel observation of a highly structured particle distribution suggests a cascadelike process in velocity space. Complex velocity space structure is investigated using a three-dimensional Hermite transform, revealing, for the first time in observational data, a power-law distribution of moments. In analogy to hydrodynamics, a Kolmogorov approach leads directly to a range of predictions for this phase-space transport. The scaling theory is found to be in agreement with observations. The combined use of state-of-the-art MMS data sets, novel implementation of a Hermite transform method, and scaling theory of the velocity cascade opens new pathways to the understanding of plasma turbulence and the crucial velocity space features that lead to dissipation in plasmas.

  16. Experimental observations of low-velocity collisional systems

    NASA Astrophysics Data System (ADS)

    Jorges, Jeffery; Dove, Adrienne; Colwell, Joshua

    Low-velocity collisions in systems of centimeter-sized objects may result in particle growth by accretion, rebounding, or erosive processes that result in the production of additional smaller particles. Numerical simulations of these systems are limited by a need to understand the collisional parameters governing the outcomes of these collisions over a range of conditions. Here, we present the results from laboratory experiments designed to explore low-velocity collisions by conducting experiments in a vacuum chamber in our 0.8-sec drop tower apparatus. These experiments utilize a variety of impacting spheres, including glass, Teflon, aluminum, stainless steel, and brass. These spheres are either used in their natural state or are ``mantled'' - coated with a few-mm thick layer of a cohesive powder. A high-speed, high-resolution video camera is used to record the motion of the colliding bodies. These videos are then processed and we track the particles to determine impactor speeds before and after collision and the collisional outcome. We determine how the coefficient of restitution varies as a function of material type, morphology, and impact velocity. For impact velocities in the range from about 20-100 cm/s we observe that mantling of particles has the most significant effect, reducing the coefficients of restitution.

  17. Predicting viscous-range velocity gradient dynamics in large-eddy simulations of turbulence

    NASA Astrophysics Data System (ADS)

    Johnson, Perry; Meneveau, Charles

    2017-11-01

    The details of small-scale turbulence are not directly accessible in large-eddy simulations (LES), posing a modeling challenge because many important micro-physical processes depend strongly on the dynamics of turbulence in the viscous range. Here, we introduce a method for coupling existing stochastic models for the Lagrangian evolution of the velocity gradient tensor with LES to simulate unresolved dynamics. The proposed approach is implemented in LES of turbulent channel flow and detailed comparisons with DNS are carried out. An application to modeling the fate of deformable, small (sub-Kolmogorov) droplets at negligible Stokes number and low volume fraction with one-way coupling is carried out. These results illustrate the ability of the proposed model to predict the influence of small scale turbulence on droplet micro-physics in the context of LES. This research was made possible by a graduate Fellowship from the National Science Foundation and by a Grant from The Gulf of Mexico Research Initiative.

  18. Initial drop size and velocity distributions for airblast coaxial atomizers

    NASA Technical Reports Server (NTRS)

    Eroglu, H.; Chigier, N.

    1991-01-01

    Phase Doppler measurements were used to determine initial drop size and velocity distributions after a complete disintegration of coaxial liquid jets. The Sauter mean diameter (SMD) distribution was found to be strongly affected by the structure and behavior of the preceding liquid intact jet. The axial measurement stations were determined from the photographs of the coaxial liquid jet at very short distances (1-2 mm) downstream of the observed break-up locations. Minimum droplet mean velocities were found at the center, and maximum velocities were near the spray boundary. Size-velocity correlations show that the velocity of larger drops did not change with drop size. Drop rms velocity distributions have double peaks whose radial positions coincide with the maximum mean velocity gradients.

  19. Monte-Carlo Method Application for Precising Meteor Velocity from TV Observations

    NASA Astrophysics Data System (ADS)

    Kozak, P.

    2014-12-01

    Monte-Carlo method (method of statistical trials) as an application for meteor observations processing was developed in author's Ph.D. thesis in 2005 and first used in his works in 2008. The idea of using the method consists in that if we generate random values of input data - equatorial coordinates of the meteor head in a sequence of TV frames - in accordance with their statistical distributions we get a possibility to plot the probability density distributions for all its kinematical parameters, and to obtain their mean values and dispersions. At that the theoretical possibility appears to precise the most important parameter - geocentric velocity of a meteor - which has the highest influence onto precision of meteor heliocentric orbit elements calculation. In classical approach the velocity vector was calculated in two stages: first we calculate the vector direction as a vector multiplication of vectors of poles of meteor trajectory big circles, calculated from two observational points. Then we calculated the absolute value of velocity independently from each observational point selecting any of them from some reasons as a final parameter. In the given method we propose to obtain a statistical distribution of velocity absolute value as an intersection of two distributions corresponding to velocity values obtained from different points. We suppose that such an approach has to substantially increase the precision of meteor velocity calculation and remove any subjective inaccuracies.

  20. Development of core ion temperature gradients and edge sheared flows in a helicon plasma device investigated by laser induced fluorescence measurements

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Thakur, S. C.; Tynan, G. R.; Center for Energy Research, University of California at San Diego, San Diego, California 92093

    2016-08-15

    We report experimental observation of ion heating and subsequent development of a prominent ion temperature gradient in the core of a linear magnetized plasma device, and the controlled shear de-correlation experiment. Simultaneously, we also observe the development of strong sheared flows at the edge of the device. Both the ion temperature and the azimuthal velocity profiles are quite flat at low magnetic fields. As the magnetic field is increased, the core ion temperature increases, producing centrally peaked ion temperature profiles and therefore strong radial gradients in the ion temperature. Similarly, we observe the development of large azimuthal flows at themore » edge, with increasing magnetic field, leading to strong radially sheared plasma flows. The ion velocities and temperatures are derived from laser induced fluorescence measurements of Doppler resolved velocity distribution functions of argon ions. These features are consistent with the previous observations of simultaneously existing radially separated multiple plasma instabilities that exhibit complex plasma dynamics in a very simple plasma system. The ion temperature gradients in the core and the radially sheared azimuthal velocities at the edge point to mechanisms that can drive the multiple plasma instabilities, that were reported earlier.« less

  1. Assessment of Early Diastolic Strain-Velocity Temporal Relationships Using SPAMM-PAV (SPAtial Modulation of Magnetization with Polarity Alternating Velocity encoding)

    PubMed Central

    Zhang, Ziheng; Dione, Donald P.; Brown, Peter B.; Shapiro, Erik M.; Sinusas, Albert J.; Sampath, Smita

    2011-01-01

    A novel MR imaging technique, spatial modulation of magnetization with polarity alternating velocity encoding (SPAMM-PAV), is presented to simultaneously examine the left ventricular early diastolic temporal relationships between myocardial deformation and intra-cavity hemodynamics with a high temporal resolution of 14 ms. This approach is initially evaluated in a dynamic flow and tissue mimicking phantom. A comparison of regional longitudinal strains and intra-cavity pressure differences (integration of computed in-plane pressure gradients within a selected region) in relation to mitral valve inflow velocities is performed in eight normal volunteers. Our results demonstrate that apical regions have higher strain rates (0.145 ± 0.005 %/ms) during the acceleration period of rapid filling compared to mid-ventricular (0.114 ± 0.007 %/ms) and basal regions (0.088 ± 0.009 %/ms), and apical strain curves plateau at peak mitral inflow velocity. This pattern is reversed during the deceleration period, when the strain-rates in the basal regions are the highest (0.027 ± 0.003 %/ms) due to ongoing basal stretching. A positive base-to-apex gradient in peak pressure difference is observed during acceleration, followed by a negative base-to apex gradient during deceleration. These studies shed insight into the regional volumetric and pressure difference changes in the left ventricle during early diastolic filling. PMID:21630348

  2. Seismic velocity and attenuation structures at the top 400 km of the inner core

    NASA Astrophysics Data System (ADS)

    Yu, W.; Wen, L.; Niu, F.

    2002-12-01

    Recent seismic studies reveal an ``east-west" hemispherical difference in seismic velocity and attenuation in the top of the inner core [Niu and Wen, 2001, Wen and Niu, 2002]. The PKiKP-PKIKP observations they used only allowed them to constrain the seismic structure in the top 80 km of the inner core. The question now arises as such to what depth this hemispherical difference persists. To answer this question, we combine the PKiKP-PKIKP dataset and the PKPbc-PKIKP observations at the distance range of 147o-160o to study seismic velocity and attenuation structures in the top 400 km of the inner core along the ``equatorial paths" (the paths whose ray angles > 35o from the polar direction). We select PKPbc-PKIKP waveforms from recordings in the Global Seismic Network (GSN) and several dense regional seismic arrays. We choose recordings for events from 1990 to 2000 with simple source time functions, so only those of intermediate and deep earthquakes are used. The observed PKPbc-PKIKP differential travel times and PKIKP/PKPbc amplitude ratios exhibit an ``east-west" hemispherical difference. The PKPbc-PKIKP travel time residuals are about 0.7 second larger for those sampling the ``eastern" hemisphere than those sampling the ``western" hemisphere. The PKIKP/PKPbc amplitude ratios are generally smaller for those sampling the ``eastern" hemisphere. We construct two seismic velocity and attenuation models, with one for each ``hemisphere", by iteratively modeling the observed PKiKP-PKIKP waveforms, the PKPbc-PKIKP differential travel times and the PKIKP/PKPbc amplitude ratios. For the ``eastern" hemisphere, the observations indicate that the E1 velocity gradient and Q structure, inferred from the PKiKP-PKIKP observations sampling the top 80 km of the inner core, extend at least to 230 km inside the inner core. A change of velocity gradient and Q value is required in the deeper portion of the inner core. For the ``western" hemisphere, on the other hand, W2 velocity gradient

  3. Uncertainty based pressure reconstruction from velocity measurement with generalized least squares

    NASA Astrophysics Data System (ADS)

    Zhang, Jiacheng; Scalo, Carlo; Vlachos, Pavlos

    2017-11-01

    A method using generalized least squares reconstruction of instantaneous pressure field from velocity measurement and velocity uncertainty is introduced and applied to both planar and volumetric flow data. Pressure gradients are computed on a staggered grid from flow acceleration. The variance-covariance matrix of the pressure gradients is evaluated from the velocity uncertainty by approximating the pressure gradient error to a linear combination of velocity errors. An overdetermined system of linear equations which relates the pressure and the computed pressure gradients is formulated and then solved using generalized least squares with the variance-covariance matrix of the pressure gradients. By comparing the reconstructed pressure field against other methods such as solving the pressure Poisson equation, the omni-directional integration, and the ordinary least squares reconstruction, generalized least squares method is found to be more robust to the noise in velocity measurement. The improvement on pressure result becomes more remarkable when the velocity measurement becomes less accurate and more heteroscedastic. The uncertainty of the reconstructed pressure field is also quantified and compared across the different methods.

  4. Contrast Gradient-Based Blood Velocimetry With Computed Tomography: Theory, Simulations, and Proof of Principle in a Dynamic Flow Phantom.

    PubMed

    Korporaal, Johannes G; Benz, Matthias R; Schindera, Sebastian T; Flohr, Thomas G; Schmidt, Bernhard

    2016-01-01

    The aim of this study was to introduce a new theoretical framework describing the relationship between the blood velocity, computed tomography (CT) acquisition velocity, and iodine contrast enhancement in CT images, and give a proof of principle of contrast gradient-based blood velocimetry with CT. The time-averaged blood velocity (v(blood)) inside an artery along the axis of rotation (z axis) is described as the mathematical division of a temporal (Hounsfield unit/second) and spatial (Hounsfield unit/centimeter) iodine contrast gradient. From this new theoretical framework, multiple strategies for calculating the time-averaged blood velocity from existing clinical CT scan protocols are derived, and contrast gradient-based blood velocimetry was introduced as a new method that can calculate v(blood) directly from contrast agent gradients and the changes therein. Exemplarily, the behavior of this new method was simulated for image acquisition with an adaptive 4-dimensional spiral mode consisting of repeated spiral acquisitions with alternating scan direction. In a dynamic flow phantom with flow velocities between 5.1 and 21.2 cm/s, the same acquisition mode was used to validate the simulations and give a proof of principle of contrast gradient-based blood velocimetry in a straight cylinder of 2.5 cm diameter, representing the aorta. In general, scanning with the direction of blood flow results in decreased and scanning against the flow in increased temporal contrast agent gradients. Velocity quantification becomes better for low blood and high acquisition speeds because the deviation of the measured contrast agent gradient from the temporal gradient will increase. In the dynamic flow phantom, a modulation of the enhancement curve, and thus alternation of the contrast agent gradients, can be observed for the adaptive 4-dimensional spiral mode and is in agreement with the simulations. The measured flow velocities in the downslopes of the enhancement curves were in good

  5. Velocity-resolved observations of water in Comet Halley

    NASA Technical Reports Server (NTRS)

    Larson, Harold P.; Davis, D. Scott; Mumma, Michael J.; Weaver, Harold A.

    1986-01-01

    High resolution (lambda/delta lambda approx. = 3 x 10 to the 5th power) near-infrared observations of H2O emission from Comet Halley were acquired at the time of maximum post-perihelion geocentric Doppler shift. The observed widths and absolute positions of the H2O line profiles reveal characteristics of the molecular velocity field in the coma. These results support H2O outflow from a Sun-lit hemisphere or the entire nucleus, but not from a single, narrow jet emanating from the nucleus. The measured pre- and post-perihelion outflow velocities were 0.9 + or - 0.2 and 1.4 + or - 0.2 km/s, respectively. Temporal variations in the kinematic properties of the outflow were inferred from changes in the spectral line shapes. These results are consistent with the release of H2O into the coma from multiple jets.

  6. Analysis of magnetic gradients to study gravitropism.

    PubMed

    Hasenstein, Karl H; John, Susan; Scherp, Peter; Povinelli, Daniel; Mopper, Susan

    2013-01-01

    Gravitropism typically is generated by dense particles that respond to gravity. Experimental stimulation by high-gradient magnetic fields provides a new approach to selectively manipulate the gravisensing system. The movement of corn, wheat, and potato starch grains in suspension was examined with videomicroscopy during parabolic flights that generated 20 to 25 s of weightlessness. During weightlessness, a magnetic gradient was generated by inserting a wedge into a uniform, external magnetic field that caused repulsion of starch grains. The resultant velocity of movement was compared with the velocity of sedimentation under 1 g conditions. The high-gradient magnetic fields repelled the starch grains and generated a force of at least 0.6 g. Different wedge shapes significantly affected starch velocity and directionality of movement. Magnetic gradients are able to move diamagnetic compounds under weightless or microgravity conditions and serve as directional stimulus during seed germination in low-gravity environments. Further work can determine whether gravity sensing is based on force or contact between amyloplasts and statocyte membrane system.

  7. Sodium Atoms in the Lunar Exotail: Observed Velocity and Spatial Distributions

    NASA Technical Reports Server (NTRS)

    Line, Michael R.; Mierkiewicz, E. J.; Oliversen, R. J.; Wilson, J. K.; Haffner, L. M.; Roesler, F. L.

    2011-01-01

    The lunar sodium tail extends long distances due to radiation pressure on sodium atoms in the lunar exosphere. Our earlier observations determined the average radial velocity of sodium atoms moving down the lunar tail beyond Earth along the Sun-Moon-Earth line (i.e., the anti-lunar point) to be 12.4 km/s. Here we use the Wisconsin H-alpha Mapper to obtain the first kinematically resolved maps of the intensity and velocity distribution of this emission over a 15 x times 15 deg region on the sky near the anti-lunar point. We present both spatially and spectrally resolved observations obtained over four nights around new moon in October 2007. The spatial distribution of the sodium atoms is elongated along the ecliptic with the location of the peak intensity drifting 3 degrees east along the ecliptic per night. Preliminary modeling results suggest that the spatial and velocity distributions in the sodium exotail are sensitive to the near surface lunar sodium velocity distribution and that observations of this sort along with detailed modeling offer new opportunities to describe the time history of lunar surface sputtering over several days.

  8. Orographic precipitation and vertical velocity characteristics from drop size and fall velocity spectra observed by disdrometers

    NASA Astrophysics Data System (ADS)

    Lee, Dong-In; Kim, Dong-Kyun; Kim, Ji-Hyeon; Kang, Yunhee; Kim, Hyeonjoon

    2017-04-01

    During a summer monsoon season each year, severe weather phenomena caused by front, mesoscale convective systems, or typhoons often occur in the southern Korean Peninsula where is mostly comprised of complex high mountains. These areas play an important role in controlling formation, amount, and distribution of rainfall. As precipitation systems move over the mountains, they can develop rapidly and produce localized heavy rainfall. Thus observational analysis in the mountainous areas is required for studying terrain effects on the rapid rainfall development and its microphysics. We performed intensive field observations using two s-band operational weather radars around Mt. Jiri (1950 m ASL) during summertime on June and July in 2015-2016. Observation data of DSD (Drop Size Distribution) from Parsivel disdrometer and (w component) vertical velocity data from ultrasonic anemometers were analyzed for Typhoon Chanhom on 12 July 2015 and the heavy rain event on 1 July 2016. During the heavy rain event, a dual-Doppler radar analysis using Jindo radar and Gunsan radar was also conducted to examine 3-D wind fields and vertical structure of reflectivity in these areas. For examining up-/downdrafts in the windward or leeward side of Mt. Jiri, we developed a new scheme technique to estimate vertical velocities (w) from drop size and fall velocity spectra of Parsivel disdrometers at different stations. Their comparison with the w values observed by the 3D anemometer showed quite good agreement each other. The Z histogram with regard to the estimated w was similar to that with regard to R, indicating that Parsivel-estimated w is quite reasonable for classifying strong and weak rain, corresponding to updraft and downdraft, respectively. Mostly, positive w values (upward) were estimated in heavy rainfall at the windward side (D1 and D2). Negative w values (downward) were dominant even during large rainfall at the leeward side (D4). For D1 and D2, the upward w percentages were

  9. Study of the Pressure and Velocity Across the Aortic Valve

    NASA Astrophysics Data System (ADS)

    Kyung, Seo Young; Chung, Erica Soyun; Lee, Joo Hee; Kyung, Hayoung; Choi, Si Young

    Biomechanics of the heart, requiring an extensive understanding of the complexity of the heart, have become the interests of many biomedical engineers in cardiology today. In order to study aortic valve disease, engineers have focused on the data obtained through bio-fluid flow analysis. To further this study, physical and computational analysis on the biomechanical determinants of blood flow in the stenosed aortic valve have been examined. These observations, along with the principles of cardiovascular physiology, confirm that when blood flows through the valve opening, pressure gradient across the valve is produced as a result of stenosis of the aortic valve. The aortic valve gradient is used to interpret the increase and decrease on each side of the defective valve. To compute different pressure gradients across the aortic valve, this paper analyzes Aortic Valve Areas (AVA) using simulations based on the continuity equation and Gorlin equation. The data obtained from such analysis consist of patients in the AS category that display mild Aortic Valve Velocity (AVV) and pressure gradient. Such correlation results in the construction of a dependent relationship between severe AS causing LV systolic dysfunction and the transaortic velocity.

  10. Observations of disk-shaped bodies in free flight at terminal velocity

    NASA Technical Reports Server (NTRS)

    Vorreiter, J. W.; Tate, D. L.

    1973-01-01

    Ten disk-shaped models of a proposed nuclear heat source module were released from an aircraft and observed by radar. The initial launch attitude, spin rate, and mass of the models were varied. Significant differences were observed in the mode of flight and terminal velocity among models of different mass and launch attitudes. The data were analyzed to yield lift and drag coefficients as a function of Reynolds number. The total sea-level velocity of the models was found to be well correlated as a function of mass per unit frontal area. The demonstrated terminal velocity of the modular disk heat source, about 27 m/sec for this specific design, is only 33% of that of existing heat source designs.

  11. Focused terahertz waves generated by a phase velocity gradient in a parallel-plate waveguide.

    PubMed

    McKinney, Robert W; Monnai, Yasuaki; Mendis, Rajind; Mittleman, Daniel

    2015-10-19

    We demonstrate the focusing of a free-space THz beam emerging from a leaky parallel-plate waveguide (PPWG). Focusing is accomplished by grading the launch angle of the leaky wave using a PPWG with gradient plate separation. Inside the PPWG, the phase velocity of the guided TE1 mode exceeds the vacuum light speed, allowing the wave to leak into free space from a slit cut along the top plate. Since the leaky wave angle changes as the plate separation decreases, the beam divergence can be controlled by grading the plate separation along the propagation axis. We experimentally demonstrate focusing of the leaky wave at a selected location at frequencies of 100 GHz and 170 GHz, and compare our measurements with numerical simulations. The proposed concept can be valuable for implementing a flat and wide-aperture beam-former for THz communications systems.

  12. Observations of the scale-dependent turbulence and evaluation of the flux–gradient relationship for sensible heat for a closed Douglas-fir canopy in very weak wind conditions

    DOE PAGES

    Vickers, D.; Thomas, C. K.

    2014-09-16

    Observations of the scale-dependent turbulent fluxes, variances, and the bulk transfer parameterization for sensible heat above, within, and beneath a tall closed Douglas-fir canopy in very weak winds are examined. The daytime sub-canopy vertical velocity spectra exhibit a double-peak structure with peaks at timescales of 0.8 s and 51.2 s. A double-peak structure is also observed in the daytime sub-canopy heat flux co-spectra. The daytime momentum flux co-spectra in the upper bole space and in the sub-canopy are characterized by a relatively large cross-wind component, likely due to the extremely light and variable winds, such that the definition of amore » mean wind direction, and subsequent partitioning of the momentum flux into along- and cross-wind components, has little physical meaning. Positive values of both momentum flux components in the sub-canopy contribute to upward transfer of momentum, consistent with the observed sub-canopy secondary wind speed maximum. For the smallest resolved scales in the canopy at nighttime, we find increasing vertical velocity variance with decreasing timescale, consistent with very small eddies possibly generated by wake shedding from the canopy elements that transport momentum, but not heat. Unusually large values of the velocity aspect ratio within the canopy were observed, consistent with enhanced suppression of the horizontal wind components compared to the vertical by the very dense canopy. The flux–gradient approach for sensible heat flux is found to be valid for the sub-canopy and above-canopy layers when considered separately in spite of the very small fluxes on the order of a few W m −2 in the sub-canopy. However, single-source approaches that ignore the canopy fail because they make the heat flux appear to be counter-gradient when in fact it is aligned with the local temperature gradient in both the sub-canopy and above-canopy layers. While sub-canopy Stanton numbers agreed well with values typically reported

  13. Observations of the scale-dependent turbulence and evaluation of the flux–gradient relationship for sensible heat for a closed Douglas-fir canopy in very weak wind conditions

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Vickers, D.; Thomas, C. K.

    Observations of the scale-dependent turbulent fluxes, variances, and the bulk transfer parameterization for sensible heat above, within, and beneath a tall closed Douglas-fir canopy in very weak winds are examined. The daytime sub-canopy vertical velocity spectra exhibit a double-peak structure with peaks at timescales of 0.8 s and 51.2 s. A double-peak structure is also observed in the daytime sub-canopy heat flux co-spectra. The daytime momentum flux co-spectra in the upper bole space and in the sub-canopy are characterized by a relatively large cross-wind component, likely due to the extremely light and variable winds, such that the definition of amore » mean wind direction, and subsequent partitioning of the momentum flux into along- and cross-wind components, has little physical meaning. Positive values of both momentum flux components in the sub-canopy contribute to upward transfer of momentum, consistent with the observed sub-canopy secondary wind speed maximum. For the smallest resolved scales in the canopy at nighttime, we find increasing vertical velocity variance with decreasing timescale, consistent with very small eddies possibly generated by wake shedding from the canopy elements that transport momentum, but not heat. Unusually large values of the velocity aspect ratio within the canopy were observed, consistent with enhanced suppression of the horizontal wind components compared to the vertical by the very dense canopy. The flux–gradient approach for sensible heat flux is found to be valid for the sub-canopy and above-canopy layers when considered separately in spite of the very small fluxes on the order of a few W m −2 in the sub-canopy. However, single-source approaches that ignore the canopy fail because they make the heat flux appear to be counter-gradient when in fact it is aligned with the local temperature gradient in both the sub-canopy and above-canopy layers. While sub-canopy Stanton numbers agreed well with values typically reported

  14. Sea surface velocities from visible and infrared multispectral atmospheric mapping sensor imagery

    NASA Technical Reports Server (NTRS)

    Pope, P. A.; Emery, W. J.; Radebaugh, M.

    1992-01-01

    High resolution (100 m), sequential Multispectral Atmospheric Mapping Sensor (MAMS) images were used in a study to calculate advective surface velocities using the Maximum Cross Correlation (MCC) technique. Radiance and brightness temperature gradient magnitude images were formed from visible (0.48 microns) and infrared (11.12 microns) image pairs, respectively, of Chandeleur Sound, which is a shallow body of water northeast of the Mississippi delta, at 145546 GMT and 170701 GMT on 30 Mar. 1989. The gradient magnitude images enhanced the surface water feature boundaries, and a lower cutoff on the gradient magnitudes calculated allowed the undesirable sunglare and backscatter gradients in the visible images, and the water vapor absorption gradients in the infrared images, to be reduced in strength. Requiring high (greater than 0.4) maximum cross correlation coefficients and spatial coherence of the vector field aided in the selection of an optimal template size of 10 x 10 pixels (first image) and search limit of 20 pixels (second image) to use in the MCC technique. Use of these optimum input parameters to the MCC algorithm, and high correlation and spatial coherence filtering of the resulting velocity field from the MCC calculation yielded a clustered velocity distribution over the visible and infrared gradient images. The velocity field calculated from the visible gradient image pair agreed well with a subjective analysis of the motion, but the velocity field from the infrared gradient image pair did not. This was attributed to the changing shapes of the gradient features, their nonuniqueness, and large displacements relative to the mean distance between them. These problems implied a lower repeat time for the imagery was needed in order to improve the velocity field derived from gradient imagery. Suggestions are given for optimizing the repeat time of sequential imagery when using the MCC method for motion studies. Applying the MCC method to the infrared

  15. Observation of Wave Packet Distortion during a Negative-Group-Velocity Transmission

    PubMed Central

    Ye, Dexin; Salamin, Yannick; Huangfu, Jiangtao; Qiao, Shan; Zheng, Guoan; Ran, Lixin

    2015-01-01

    In Physics, causality is a fundamental postulation arising from the second law of thermodynamics. It states that, the cause of an event precedes its effect. In the context of Electromagnetics, the relativistic causality limits the upper bound of the velocity of information, which is carried by electromagnetic wave packets, to the speed of light in free space (c). In anomalously dispersive media (ADM), it has been shown that, wave packets appear to propagate with a superluminal or even negative group velocity. However, Sommerfeld and Brillouin pointed out that the “front” of such wave packets, known as the initial point of the Sommerfeld precursor, always travels at c. In this work, we investigate the negative-group-velocity transmission of half-sine wave packets. We experimentally observe the wave front and the distortion of modulated wave packets propagating with a negative group velocity in a passive artificial ADM in microwave regime. Different from previous literature on the propagation of superluminal Gaussian packets, strongly distorted sinusoidal packets with non-superluminal wave fronts were observed. This result agrees with Brillouin's assertion, i.e., the severe distortion of seemingly superluminal wave packets makes the definition of group velocity physically meaningless in the anomalously dispersive region. PMID:25631746

  16. On the impact of adverse pressure gradient on the supersonic turbulent boundary layer

    NASA Astrophysics Data System (ADS)

    Wang, Qian-Cheng; Wang, Zhen-Guo; Zhao, Yu-Xin

    2016-11-01

    By employing the particle image velocimetry, the mean and turbulent characteristics of a Mach 2.95 turbulent boundary layer are experimentally investigated without the impact of curvature. The physical mechanism with which the streamwise adverse pressure gradient affects the supersonic boundary layer is revealed. The data are compared to that of the concave boundary layer with similar streamwise distributions of wall static pressure to clarify the separate impacts of the adverse pressure gradient and the concave curvature. The logarithmic law is observed to be well preserved for both of the cases. The dip below the logarithmic law is not observed in present investigation. Theoretical analysis indicates that it could be the result of compromise between the opposite impacts of the compression wave and the increased turbulent intensity. Compared to the zero pressure gradient boundary layer, the principal strain rate and the turbulent intensities are increased by the adverse pressure gradient. The shear layer formed due the hairpin packets could be sharpened by the compression wave, which leads to higher principal strain rate and the associated turbulent level. Due to the additional impact of the centrifugal instability brought by the concave wall, even higher turbulent intensities than that of the adverse pressure gradient case are introduced. The existence of velocity modes within the zero pressure gradient boundary layer suggests that the large scale motions are statistically well organized. The generation of new velocity modes due to the adverse pressure gradient indicates that the turbulent structure is changed by the adverse pressure gradient, through which more turbulence production that cannot be effectively predicted by the Reynolds-stress transport equations could be brought.

  17. Effect of a surface tension gradient on the slip flow along a superhydrophobic air-water interface

    NASA Astrophysics Data System (ADS)

    Song, Dong; Song, Baowei; Hu, Haibao; Du, Xiaosong; Du, Peng; Choi, Chang-Hwan; Rothstein, Jonathan P.

    2018-03-01

    Superhydrophobic surfaces have been shown to produce significant drag reduction in both laminar and turbulent flows by introducing an apparent slip velocity along an air-water interface trapped within the surface roughness. In the experiments presented within this study, we demonstrate the existence of a surface tension gradient associated with the resultant Marangoni flow along an air-water interface that causes the slip velocity and slip length to be significantly reduced. In this study, the slip velocity along a millimeter-sized air-water interface was investigated experimentally. This large-scale air-water interface facilitated a detailed investigation of the interfacial velocity profiles as the flow rate, interfacial curvature, and interface geometry were varied. For the air-water interfaces supported above continuous grooves (concentric rings within a torsional shear flow) where no surface tension gradient exists, a slip velocity as high as 30% of the bulk velocity was observed. However, for the air-water interfaces supported above discontinuous grooves (rectangular channels in a Poiseuille flow), the presence of a surface tension gradient reduced the slip velocity and in some cases resulted in an interfacial velocity that was opposite to the main flow direction. The curvature of the air-water interface in the spanwise direction was found to dictate the details of the interfacial flow profile with reverse flow in the center of the interface for concave surfaces and along the outside of the interface for convex surfaces. The deflection of the air-water interface was also found to greatly affect the magnitude of the slip. Numerical simulations imposed with a relatively small surface tension gradient along the air-water interface were able to predict both the reduced slip velocity and back flow along the air-water interface.

  18. Measurements of phoretic velocities of aerosol particles in microgravity conditions

    NASA Astrophysics Data System (ADS)

    Prodi, F.; Santachiara, G.; Travaini, S.; Vedernikov, A.; Dubois, F.; Minetti, C.; Legros, J. C.

    2006-11-01

    Measurements of thermo- and diffusio-phoretic velocities of aerosol particles (carnauba wax, paraffin and sodium chloride) were performed in microgravity conditions (Drop Tower facility, in Bremen, and Parabolic Flights, in Bordeaux). In the case of thermophoresis, a temperature gradient was obtained by heating the upper plate of the cell, while the lower one was maintained at environmental temperature. For diffusiophoresis, the water vapour gradient was obtained with sintered plates imbued with a water solution of MgCl 2 and distilled water, at the top and at the bottom of the cell, respectively. Aerosol particles were observed through a digital holographic velocimeter, a device allowing the determination of 3-D coordinates of particles from the observed volume. Particle trajectories and consequently particle velocities were reconstructed through the analysis of the sequence of particle positions. The experimental values of reduced thermophoretic velocities are between the theoretical values of Yamamoto and Ishihara [Yamamoto, K., Ishihara, Y., 1988. Thermophoresis of a spherical particle in a rarefied gas of a transition regime. Phys. Fluids. 31, 3618-3624] and Talbot et al. [Talbot, L., Cheng, R.K., Schefer, R.W., Willis, D.R., 1980. Thermophoresis of particles in a heated boundary layer. J. Fluid Mech. 101, 737-758], and do not show a clear dependence on the thermal conductivity of the aerosol. The existence of negative thermophoresis is not confirmed in our experiments. Concerning diffusiophoretic experiments, the results obtained show a small increase of reduced diffusiophoretic velocity with the Knudsen number.

  19. Large-Scale periodic solar velocities: An observational study

    NASA Technical Reports Server (NTRS)

    Dittmer, P. H.

    1977-01-01

    Observations of large-scale solar velocities were made using the mean field telescope and Babcock magnetograph of the Stanford Solar Observatory. Observations were made in the magnetically insensitive ion line at 5124 A, with light from the center (limb) of the disk right (left) circularly polarized, so that the magnetograph measures the difference in wavelength between center and limb. Computer calculations are made of the wavelength difference produced by global pulsations for spherical harmonics up to second order and of the signal produced by displacing the solar image relative to polarizing optics or diffraction grating.

  20. Copernicus observations of Iota Herculis velocity variations

    NASA Technical Reports Server (NTRS)

    Rogerson, J. B., Jr.

    1984-01-01

    Observations of Iota Her at 109.61-109.67 nm obtained with the U1 channel of the Copernicus spectrophotometer at resolution 5 pm during 3.6 days in May, 1979, are reported. Radial-velocity variations are detected and analyzed as the sum of two sinusoids with frequencies 0.660 and 0.618 cycles/day and amplitudes 9.18 and 8.11 km/s, respectively. Weak evidence supporting the 13.9-h periodicity seen in line-profile variations by Smith (1978) is found.

  1. Radar observations of density gradients, electric fields, and plasma irregularities near polar cap patches in the context of the gradient-drift instability

    NASA Astrophysics Data System (ADS)

    Lamarche, Leslie J.; Makarevich, Roman A.

    2017-03-01

    We present observations of plasma density gradients, electric fields, and small-scale plasma irregularities near a polar cap patch made by the Super Dual Auroral Radar Network radar at Rankin Inlet (RKN) and the northern face of Resolute Bay Incoherent Scatter Radar (RISR-N). RKN echo power and occurrence are analyzed in the context of gradient-drift instability (GDI) theory, with a particular focus on the previously uninvestigated 2-D dependencies on wave propagation, electric field, and gradient vectors, with the latter two quantities evaluated directly from RISR-N measurements. It is shown that higher gradient and electric field components along the wave vector generally lead to the higher observed echo occurrence, which is consistent with the expected higher GDI growth rate, but the relationship with echo power is far less straightforward. The RKN echo power increases monotonically as the predicted linear growth rate approaches zero from negative values but does not continue this trend into positive growth rate values, in contrast with GDI predictions. The observed greater consistency of echo occurrence with GDI predictions suggests that GDI operating in the linear regime can control basic plasma structuring, but measured echo strength may be affected by other processes and factors, such as multistep or nonlinear processes or a shear-driven instability.

  2. Field-gradient partitioning for fracture and frictional contact in the material point method: Field-gradient partitioning for fracture and frictional contact in the material point method [Fracture and frictional contact in material point method using damage-field gradients for velocity-field partitioning

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Homel, Michael A.; Herbold, Eric B.

    Contact and fracture in the material point method require grid-scale enrichment or partitioning of material into distinct velocity fields to allow for displacement or velocity discontinuities at a material interface. We present a new method which a kernel-based damage field is constructed from the particle data. The gradient of this field is used to dynamically repartition the material into contact pairs at each node. Our approach avoids the need to construct and evolve explicit cracks or contact surfaces and is therefore well suited to problems involving complex 3-D fracture with crack branching and coalescence. A straightforward extension of this approachmore » permits frictional ‘self-contact’ between surfaces that are initially part of a single velocity field, enabling more accurate simulation of granular flow, porous compaction, fragmentation, and comminution of brittle materials. Finally, numerical simulations of self contact and dynamic crack propagation are presented to demonstrate the accuracy of the approach.« less

  3. Field-gradient partitioning for fracture and frictional contact in the material point method: Field-gradient partitioning for fracture and frictional contact in the material point method [Fracture and frictional contact in material point method using damage-field gradients for velocity-field partitioning

    DOE PAGES

    Homel, Michael A.; Herbold, Eric B.

    2016-08-15

    Contact and fracture in the material point method require grid-scale enrichment or partitioning of material into distinct velocity fields to allow for displacement or velocity discontinuities at a material interface. We present a new method which a kernel-based damage field is constructed from the particle data. The gradient of this field is used to dynamically repartition the material into contact pairs at each node. Our approach avoids the need to construct and evolve explicit cracks or contact surfaces and is therefore well suited to problems involving complex 3-D fracture with crack branching and coalescence. A straightforward extension of this approachmore » permits frictional ‘self-contact’ between surfaces that are initially part of a single velocity field, enabling more accurate simulation of granular flow, porous compaction, fragmentation, and comminution of brittle materials. Finally, numerical simulations of self contact and dynamic crack propagation are presented to demonstrate the accuracy of the approach.« less

  4. Experimental observation of electron-temperature-gradient turbulence in a laboratory plasma.

    PubMed

    Mattoo, S K; Singh, S K; Awasthi, L M; Singh, R; Kaw, P K

    2012-06-22

    We report the observation of electron-temperature-gradient (ETG) driven turbulence in the laboratory plasma of a large volume plasma device. The removal of unutilized primary ionizing and nonthermal electrons from uniform density plasma and the imposition and control of the gradient in the electron temperature (T[Symbol: see text] T(e)) are all achieved by placing a large (2 m diameter) magnetic electron energy filter in the middle of the device. In the dressed plasma, the observed ETG turbulence in the lower hybrid range of frequencies ν = (1-80 kHz) is characterized by a broadband with a power law. The mean wave number k perpendicular ρ(e) = (0.1-0.2) satisfies the condition k perpendicular ρ(e) ≤ 1, where ρ(e) is the electron Larmor radius.

  5. Uncertainty assessment of 3D instantaneous velocity model from stack velocities

    NASA Astrophysics Data System (ADS)

    Emanuele Maesano, Francesco; D'Ambrogi, Chiara

    2015-04-01

    3D modelling is a powerful tool that is experiencing increasing applications in data analysis and dissemination. At the same time the need of quantitative uncertainty evaluation is strongly requested in many aspects of the geological sciences and by the stakeholders. In many cases the starting point for 3D model building is the interpretation of seismic profiles that provide indirect information about the geology of the subsurface in the domain of time. The most problematic step in the 3D modelling construction is the conversion of the horizons and faults interpreted in time domain to the depth domain. In this step the dominant variable that could lead to significantly different results is the velocity. The knowledge of the subsurface velocities is related mainly to punctual data (sonic logs) that are often sparsely distributed in the areas covered by the seismic interpretation. The extrapolation of velocity information to wide extended horizons is thus a critical step to obtain a 3D model in depth that can be used for predictive purpose. In the EU-funded GeoMol Project, the availability of a dense network of seismic lines (confidentially provided by ENI S.p.A.) in the Central Po Plain, is paired with the presence of 136 well logs, but few of them have sonic logs and in some portion of the area the wells are very widely spaced. The depth conversion of the 3D model in time domain has been performed testing different strategies for the use and the interpolation of velocity data. The final model has been obtained using a 4 layer cake 3D instantaneous velocity model that considers both the initial velocity (v0) in every reference horizon and the gradient of velocity variation with depth (k). Using this method it is possible to consider the geological constraint given by the geometries of the horizons and the geo-statistical approach to the interpolation of velocities and gradient. Here we present an experiment based on the use of set of pseudo-wells obtained from the

  6. Vorticity and helicity decompositions and dynamics with real Schur form of the velocity gradient

    NASA Astrophysics Data System (ADS)

    Zhu, Jian-Zhou

    2018-03-01

    The real Schur form (RSF) of a generic velocity gradient field ∇u is exploited to expose the structures of flows, in particular, our field decomposition resulting in two vorticities with only mutual linkage as the topological content of the global helicity (accordingly decomposed into two equal parts). The local transformation to the RSF may indicate alternative (co)rotating frame(s) for specifying the objective argument(s) of the constitutive equation. When ∇u is uniformly of RSF in a fixed Cartesian coordinate frame, i.e., ux = ux(x, y) and uy = uy(x, y), but uz = uz(x, y, z), the model, with the decomposed vorticities both frozen-in to u, is for two-component-two-dimensional-coupled-with-one-component-three-dimensional flows in between two-dimensional-three-component (2D3C) and fully three-dimensional-three-component ones and may help curing the pathology in the helical 2D3C absolute equilibrium, making the latter effectively work in more realistic situations.

  7. On the evolution of the invariants of the velocity gradient tensor in single-square-grid-generated turbulence

    NASA Astrophysics Data System (ADS)

    Zhou, Yi; Nagata, Koji; Sakai, Yasuhiko; Ito, Yasumasa; Hayase, Toshiyuki

    2015-07-01

    Direct numerical simulations were performed to investigate the topological evolution of turbulence generated by a single square grid. Immediately behind the single square grid (i.e., in the irrotational dissipation region), the conditional mean trajectories (CMTs) of R and Q are distinctly different from those in homogeneous isotropic turbulence (HIT), where R and Q are the third and second invariants, respectively, of the velocity gradient tensor. In this region, the non-local influence of the pressure Hessian is dominant, which causes irrotational viscous dissipation. The anisotropic part of the pressure Hessian may be responsible for the irrotational viscous dissipation found at the turbulent/nonturbulent interface in turbulent jets [C. B. da Silva and J. C. F. Pereira, "Invariants of the velocity-gradient, rate-of-strain, and rate-of-rotation tensors across the turbulent/nonturbulent interface in jets," Phys. Fluids 20, 055101 (2008) and Watanabe et al., "Vortex stretching and compression near the turbulent/non-turbulent interface in a planar jet," J. Fluid Mech. 758, 754 (2014)]. In the transition region, the CMTs of R and Q gradually acquire an evolution pattern similar to that in HIT. The expansion of the (R, Q) map at Q > 0 is associated with the effects of the restricted Euler term. Finally, in the fully turbulent region, the CMTs of R and Q demonstrate a clockwise evolution toward a point close to the origin. However, the cyclic spiraling seen in HIT is not found. The lack of the cyclic evolution may be attributed to the considerably large effect of the viscous term owing to the relatively small local Reynolds number. On average, the combined influences of the restricted Euler term and anisotropic part of the pressure Hessian contribute to the generation of small-scale motions, and the viscous term tends to remove small-scale motions.

  8. What Do the Hitomi Observations Tell Us About the Turbulent Velocities in the Perseus Cluster?

    NASA Astrophysics Data System (ADS)

    ZuHone, John A.; Miller, Eric D.; Bulbul, Esra; Zhuravleva, Irina

    2017-08-01

    Recently, the Hitomi X-ray Observatory provided the first-ever direct measurements of Doppler line shifting and broadening from the hot plasma in clusters of galaxies via its observations of the Perseus Cluster. It has been reported that these observations demonstrate that the ICM in Perseus is "quiescent". It is indisputable that the velocities inferred from the measured line shifts and broadening are low, but what do these observations imply about the structure of the velocity field on scales smaller than the Hitomi PSF? We use hydrodynamic simulations of gas motions in a cool-core cluster in combination with synthetic Hitomi observations in order to compare the observed line-of-sight velocities to the 3D velocity structure of the ICM, and assess the impact of Hitomi's spatial resolution and the effects of varying the underlying ICM physics.

  9. 2D He+ Pickup Ion Velocity Distribution Functions: STEREO PLASTIC Observations

    NASA Astrophysics Data System (ADS)

    Drews, C.; Berger, L.; Peleikis, T.; Wimmer-Schweingruber, R. F.

    2014-12-01

    He+ pickup ions are either born from the ionization of interstellar neutral helium atoms inside our heliosphere, the so called interstellar pickup ions, or through the interaction of solar wind ions with small dust particles close to the Sun, the so called inner-source of pickup ions. Until now, most observations of He+ pickup ions were limited to reduced 1D velocity spectra, which are insufficient to study certain characteristics of the He+ Velocity Distribution Function (VDF). It is generally assumed that rapid pitch-angle scattering of freshly created pickup ions quickly leads to a fully isotropic He+ VDF. In the light of recent observations, this assumption has found to be oversimplified and needs to be re-investigated. Using He+ pickup ion data from the PLASTIC instrument on board the STEREO A spacecraft we reconstruct a reduced form of the He+ VDF in 2 dimensions (see figure). The reduced form of the He+ VDF allows us to study the pitch-angle distribution and anisotropy of the He+ VDF as a function of the solar magnetic field, B. Our observations show clear signs of a significant anisotropy of the He+ VDF and even indicates that, at least for certain configurations of B, it is not even fully gyrotropic. Our results further suggest, that the observed velocity and pitch-angle of He+ depends strongly on the solar magnetic field vector, B, the ecliptic longitude, λ, the solar wind speed, vsw, and the history of B. Consequently, we argue that reduced 1D velocity spectra of He+ are insufficient to study quantities like the pitch-angle scattering rate, τ, or the adiabatic cooling index γ.

  10. The 2011 Tohoku-oki Earthquake related to a large velocity gradient within the Pacific plate

    NASA Astrophysics Data System (ADS)

    Matsubara, Makoto; Obara, Kazushige

    2015-04-01

    rays from the hypocenter around the coseismic region of the Tohoku-oki earthquake take off downward and pass through the Pacific plate. The landward low-V zone with a large anomaly corresponds to the western edge of the coseismic slip zone of the 2011 Tohoku-oki earthquake. The initial break point (hypocenter) is associated with the edge of a slightly low-V and low-Vp/Vs zone corresponding to the boundary of the low- and high-V zone. The trenchward low-V and low-Vp/Vs zone extending southwestward from the hypocenter may indicate the existence of a subducted seamount. The high-V zone and low-Vp/Vs zone might have accumulated the strain and resulted in the huge coseismic slip zone of the 2011 Tohoku earthquake. The low-V and low-Vp/Vs zone is a slight fluctuation within the high-V zone and might have acted as the initial break point of the 2011 Tohoku earthquake. Reference Matsubara, M. and K. Obara (2011) The 2011 Off the Pacific Coast of Tohoku earthquake related to a strong velocity gradient with the Pacific plate, Earth Planets Space, 63, 663-667. Okada, Y., K. Kasahara, S. Hori, K. Obara, S. Sekiguchi, H. Fujiwara, and A. Yamamoto (2004) Recent progress of seismic observation networks in Japan-Hi-net, F-net, K-NET and KiK-net, Research News Earth Planets Space, 56, xv-xxviii.

  11. DNS of a non-equilibrium adverse pressure gradient turbulent boundary layer

    NASA Astrophysics Data System (ADS)

    Gungor, Taygun R.; Gungor, Ayse G.; Maciel, Yvan; Simens, Mark P.

    2017-11-01

    A new direct numerical simulation (DNS) dataset of a non-equilibrium adverse pressure gradient (APG) turbulent boundary layer (TBL) that evolves from a zero-pressure-gradient (ZPG) TBL to a TBL which is very close to separation at Reθ is around 8200 is presented. There are two simulations running together in the DNS computational setup. The APG TBL spans Reθ = 1476 - 8276 . Mean velocity results do not satisfy the log law as the defect in the velocity increases. The production and the Reynolds stress peak are observed around y /δ* = 1 after the flow is evolved up to a certain point. The new dataset is compared with other datasets in terms of mean values, Reynolds stresses and turbulent kinetic energy budgets and using this comparison scaling study is performed. Funded by in part by ITU-AYP and NSERC of Canada.

  12. Evaluation of multiple tracer methods to estimate low groundwater flow velocities

    DOE PAGES

    Reimus, Paul W.; Arnold, Bill W.

    2017-02-20

    Here, four different tracer methods were used to estimate groundwater flow velocity at a multiple-well site in the saturated alluvium south of Yucca Mountain, Nevada: (1) two single-well tracer tests with different rest or “shut-in” periods, (2) a cross-hole tracer test with an extended flow interruption, (3) a comparison of two tracer decay curves in an injection borehole with and without pumping of a downgradient well, and (4) a natural-gradient tracer test. Such tracer methods are potentially very useful for estimating groundwater velocities when hydraulic gradients are flat (and hence uncertain) and also when water level and hydraulic conductivity datamore » are sparse, both of which were the case at this test location. The purpose of the study was to evaluate the first three methods for their ability to provide reasonable estimates of relatively low groundwater flow velocities in such low-hydraulic-gradient environments. The natural-gradient method is generally considered to be the most robust and direct method, so it was used to provide a “ground truth” velocity estimate. However, this method usually requires several wells, so it is often not practical in systems with large depths to groundwater and correspondingly high well installation costs. The fact that a successful natural gradient test was conducted at the test location offered a unique opportunity to compare the flow velocity estimates obtained by the more easily deployed and lower risk methods with the ground-truth natural-gradient method. The groundwater flow velocity estimates from the four methods agreed very well with each other, suggesting that the first three methods all provided reasonably good estimates of groundwater flow velocity at the site. We discuss the advantages and disadvantages of the different methods, as well as some of the uncertainties associated with them.« less

  13. Evaluation of multiple tracer methods to estimate low groundwater flow velocities

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Reimus, Paul W.; Arnold, Bill W.

    Here, four different tracer methods were used to estimate groundwater flow velocity at a multiple-well site in the saturated alluvium south of Yucca Mountain, Nevada: (1) two single-well tracer tests with different rest or “shut-in” periods, (2) a cross-hole tracer test with an extended flow interruption, (3) a comparison of two tracer decay curves in an injection borehole with and without pumping of a downgradient well, and (4) a natural-gradient tracer test. Such tracer methods are potentially very useful for estimating groundwater velocities when hydraulic gradients are flat (and hence uncertain) and also when water level and hydraulic conductivity datamore » are sparse, both of which were the case at this test location. The purpose of the study was to evaluate the first three methods for their ability to provide reasonable estimates of relatively low groundwater flow velocities in such low-hydraulic-gradient environments. The natural-gradient method is generally considered to be the most robust and direct method, so it was used to provide a “ground truth” velocity estimate. However, this method usually requires several wells, so it is often not practical in systems with large depths to groundwater and correspondingly high well installation costs. The fact that a successful natural gradient test was conducted at the test location offered a unique opportunity to compare the flow velocity estimates obtained by the more easily deployed and lower risk methods with the ground-truth natural-gradient method. The groundwater flow velocity estimates from the four methods agreed very well with each other, suggesting that the first three methods all provided reasonably good estimates of groundwater flow velocity at the site. We discuss the advantages and disadvantages of the different methods, as well as some of the uncertainties associated with them.« less

  14. Evaluation of multiple tracer methods to estimate low groundwater flow velocities.

    PubMed

    Reimus, Paul W; Arnold, Bill W

    2017-04-01

    Four different tracer methods were used to estimate groundwater flow velocity at a multiple-well site in the saturated alluvium south of Yucca Mountain, Nevada: (1) two single-well tracer tests with different rest or "shut-in" periods, (2) a cross-hole tracer test with an extended flow interruption, (3) a comparison of two tracer decay curves in an injection borehole with and without pumping of a downgradient well, and (4) a natural-gradient tracer test. Such tracer methods are potentially very useful for estimating groundwater velocities when hydraulic gradients are flat (and hence uncertain) and also when water level and hydraulic conductivity data are sparse, both of which were the case at this test location. The purpose of the study was to evaluate the first three methods for their ability to provide reasonable estimates of relatively low groundwater flow velocities in such low-hydraulic-gradient environments. The natural-gradient method is generally considered to be the most robust and direct method, so it was used to provide a "ground truth" velocity estimate. However, this method usually requires several wells, so it is often not practical in systems with large depths to groundwater and correspondingly high well installation costs. The fact that a successful natural gradient test was conducted at the test location offered a unique opportunity to compare the flow velocity estimates obtained by the more easily deployed and lower risk methods with the ground-truth natural-gradient method. The groundwater flow velocity estimates from the four methods agreed very well with each other, suggesting that the first three methods all provided reasonably good estimates of groundwater flow velocity at the site. The advantages and disadvantages of the different methods, as well as some of the uncertainties associated with them are discussed. Published by Elsevier B.V.

  15. Gravitation is a Gradient in the Velocity of Light

    NASA Astrophysics Data System (ADS)

    Froedge, Dt

    2017-01-01

    It is well known that a photon moving in a gravitational field has a trajectory that can be defined by Fermat's principle with a variable speed of light and no other gravitational influence. If it can be shown that a particle composed of speed of light sub-particles has the same acceleration in a variable index of refraction as a particle in a gravitational field, then there is no need to ascribe any other mechanism to gravitation than a gradient in c. This makes gravitation an electromagnetic phenomenon, and if QFT can illustrate a gradient in c can be produced by the internal motion of lightspeed sub-particles then the unification of QM and gravitation becomes more straightforward. http://www.arxdtf.org/css/GravAPS.pdf.

  16. Aortic-Brachial Pulse Wave Velocity Ratio: A Measure of Arterial Stiffness Gradient Not Affected by Mean Arterial Pressure.

    PubMed

    Fortier, Catherine; Desjardins, Marie-Pier; Agharazii, Mohsen

    2018-03-01

    Aortic stiffness, measured by carotid-femoral pulse wave velocity (cf-PWV), is used for the prediction of cardiovascular risk. This mini-review describes the nonlinear relationship between cf-PWV and operational blood pressure, presents the proposed methods to adjust for this relationship, and discusses a potential place for aortic-brachial PWV ratio (a measure of arterial stiffness gradient) as a blood pressure-independent measure of vascular aging. PWV is inherently dependent on the operational blood pressure. In cross-sectional studies, PWV adjustment for mean arterial pressure (MAP) is preferred, but still remains a nonoptimal approach, as the relationship between PWV and blood pressure is nonlinear and varies considerably among individuals due to heterogeneity in genetic background, vascular tone, and vascular remodeling. Extrapolations from the blood pressure-independent stiffness parameter β (β 0 ) have led to the creation of stiffness index β, which can be used for local stiffness. A similar approach has been used for cardio-ankle PWV to generate a blood pressure-independent cardio-ankle vascular index (CAVI). It was recently demonstrated that stiffness index β and CAVI remain slightly blood pressure-dependent, and a more appropriate formula has been proposed to make the proper adjustments. On the other hand, the negative impact of aortic stiffness on clinical outcomes is thought to be mediated through attenuation or reversal of the arterial stiffness gradient, which can also be influenced by a reduction in peripheral medium-sized muscular arteries in conditions that predispose to accelerate vascular aging. Arterial stiffness gradient, assessed by aortic-brachial PWV ratio, is emerging to be at least as good as cf-PWV for risk prediction, but has the advantage of not being affected by operating MAP. The negative impacts of aortic stiffness on clinical outcomes are proposed to be mediated through attenuation or reversal of arterial stiffness gradient

  17. Newly velocity field of Sulawesi Island from GPS observation

    NASA Astrophysics Data System (ADS)

    Sarsito, D. A.; Susilo, Simons, W. J. F.; Abidin, H. Z.; Sapiie, B.; Triyoso, W.; Andreas, H.

    2017-07-01

    Sulawesi microplate Island is located at famous triple junction area of the Eurasian, India-Australian, and Philippine Sea plates. Under the influence of the northward moving Australian plate and the westward motion of the Philippine plate, the island at Eastern part of Indonesia is collide and with the Eurasian plate and Sunda Block. Those recent microplate tectonic motions can be quantitatively determine by GNSS-GPS measurement. We use combine GNSS-GPS observation types (campaign type and continuous type) from 1997 to 2015 to derive newly velocity field of the area. Several strategies are applied and tested to get the optimum result, and finally we choose regional strategy to reduce error propagation contribution from global multi baseline processing using GAMIT/GLOBK 10.5. Velocity field are analyzed in global reference frame ITRF 2008 and local reference frame by fixing with respect alternatively to Eurasian plate - Sunda block, India-Australian plate and Philippine Sea plates. Newly results show dense distribution of velocity field. This information is useful for tectonic deformation studying in geospatial era.

  18. Velocity structure in long period variable star atmospheres

    NASA Technical Reports Server (NTRS)

    Pilachowski, C.; Wallerstein, G.; Willson, L. A.

    1980-01-01

    A regression analysis of the dependence of absorption line velocities on wavelength, line strength, excitation potential, and ionization potential is presented. The method determines the region of formation of the absorption lines for a given data and wavelength region. It is concluded that the scatter which is frequently found in velocity measurements of absorption lines in long period variables is probably the result of a shock of moderate amplitude located in or near the reversing layer and that the frequently observed correlation of velocity with excitation and ionization are a result of the velocity gradients produced by this shock in the atmosphere. A simple interpretation of the signs of the coefficients of the regression analysis is presented in terms of preshock, post shock, or across the shock, together with criteria for evaluating the validity of the fit. The amplitude of the reversing layer shock is estimated from an analysis of a series of plates for four long period variable stars along with the most probable stellar velocity for these stars.

  19. Estimating vertical velocity and radial flow from Doppler radar observations of tropical cyclones

    NASA Astrophysics Data System (ADS)

    Lee, J. L.; Lee, W. C.; MacDonald, A. E.

    2006-01-01

    The mesoscale vorticity method (MVM) is used in conjunction with the ground-based velocity track display (GBVTD) to derive the inner-core vertical velocity from Doppler radar observations of tropical cyclone (TC) Danny (1997). MVM derives the vertical velocity from vorticity variations in space and in time based on the mesoscale vorticity equation. The use of MVM and GBVTD allows us to derive good correlations among the eye-wall maximum wind, bow-shaped updraught and echo east of the eye-wall in Danny. Furthermore, we demonstrate the dynamically consistent radial flow can be derived from the vertical velocity obtained from MVM using the wind decomposition technique that solves the Poisson equations over a limited-area domain. With the wind decomposition, we combine the rotational wind which is obtained from Doppler radar wind observations and the divergent wind which is inferred dynamically from the rotational wind to form the balanced horizontal wind in TC inner cores, where rotational wind dominates the divergent wind. In this study, we show a realistic horizontal and vertical structure of the vertical velocity and the induced radial flow in Danny's inner core. In the horizontal, the main eye-wall updraught draws in significant surrounding air, converging at the strongest echo where the maximum updraught is located. In the vertical, the main updraught tilts vertically outwards, corresponding very well with the outward-tilting eye-wall. The maximum updraught is located at the inner edge of the eye-wall clouds, while downward motions are found at the outer edge. This study demonstrates that the mesoscale vorticity method can use high-temporal-resolution data observed by Doppler radars to derive realistic vertical velocity and the radial flow of TCs. The vorticity temporal variations crucial to the accuracy of the vorticity method have to be derived from a high-temporal-frequency observing system such as state-of-the-art Doppler radars.

  20. Observation of Brownian motion in liquids at short times: instantaneous velocity and memory loss.

    PubMed

    Kheifets, Simon; Simha, Akarsh; Melin, Kevin; Li, Tongcang; Raizen, Mark G

    2014-03-28

    Measurement of the instantaneous velocity of Brownian motion of suspended particles in liquid probes the microscopic foundations of statistical mechanics in soft condensed matter. However, instantaneous velocity has eluded experimental observation for more than a century since Einstein's prediction of the small length and time scales involved. We report shot-noise-limited, high-bandwidth measurements of Brownian motion of micrometer-sized beads suspended in water and acetone by an optical tweezer. We observe the hydrodynamic instantaneous velocity of Brownian motion in a liquid, which follows a modified energy equipartition theorem that accounts for the kinetic energy of the fluid displaced by the moving bead. We also observe an anticorrelated thermal force, which is conventionally assumed to be uncorrelated.

  1. Atmospheric gradients from very long baseline interferometry observations

    NASA Technical Reports Server (NTRS)

    Macmillan, D. S.

    1995-01-01

    Azimuthal asymmetries in the atmospheric refractive index can lead to errors in estimated vertical and horizontal station coordinates. Daily average gradient effects can be as large as 50 mm of delay at a 7 deg elevation. To model gradients, the constrained estimation of gradient paramters was added to the standard VLBI solution procedure. Here the analysis of two sets of data is summarized: the set of all geodetic VLBI experiments from 1990-1993 and a series of 12 state-of-the-art R&D experiments run on consecutive days in January 1994. In both cases, when the gradient parameters are estimated, the overall fit of the geodetic solution is improved at greater than the 99% confidence level. Repeatabilities of baseline lengths ranging up to 11,000 km are improved by 1 to 8 mm in a root-sum-square sense. This varies from about 20% to 40% of the total baseline length scatter without gradient modeling for the 1990-1993 series and 40% to 50% for the January series. Gradients estimated independently for each day as a piecewise linear function are mostly continuous from day to day within their formal uncertainties.

  2. Thermal-gradient migration of brine inclusions in salt crystals. [Synthetic single crystals of NaCl and KCl

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Yagnik, S.K.

    1982-09-01

    It has been proposed that high-level nuclear waste be disposed in a geologic repository. Natural-salt deposits, which are being considered for this purpose, contain a small volume fraction of water in the form of brine inclusions distributed throughout the salt. Radioactive-decay heating of the nuclear wastes will impose a temperature gradient on the surrounding salt which mobilizes the brine inclusions. Inclusions filled completely with brine migrate up the temperature gradient and eventually accumulate brine near the buried waste forms. The brine may slowly corrode or degrade the waste forms which is undesirable. In this work, thermal gradient migration of bothmore » all-liquid and gas-liquid inclusions was experimentally studied in synthetic single crystals of NaCl and KCl using a hot-stage attachment to an optical microscope which was capable of imposing temperature gradients and axial compressive loads on the crystals. The migration velocities of the inclusions were found to be dependent on temperature, temperature gradient, and inclusion shape and size. The velocities were also dictated by the interfacial mass transfer resistance at brine/solid interface. This interfacial resistance depends on the dislocation density in the crystal, which in turn, depends on the axial compressive loading of the crystal. At low axial loads, the dependence between the velocity and temperature gradient is non-linear.At high axial loads, however, the interfacial resistance is reduced and the migration velocity depends linearly on the temperature gradient. All-liquid inclusions filled with mixed brines were also studied. For gas-liquid inclusions, three different gas phases (helium, air and argon) were compared. Migration studies were also conducted on single crystallites of natural salt as well as in polycrystalline natural salt samples. The behavior of the inclusions at large angle grain boundaries was observed. 35 figures, 3 tables.« less

  3. New Synthesis of Ocean Crust Velocity Structure From Two-Dimensional Profiles

    NASA Astrophysics Data System (ADS)

    Christeson, G. L.; Goff, J.; Carlson, R. L.; Reece, R.

    2017-12-01

    The velocity structure of typical oceanic crust consists of Layer 2, where velocities increase rapidly with depth from seafloor, and Layer 3, which is thicker and has a lower velocity gradient. Previous syntheses have found no correlation of velocity structure with spreading rate, even though we know that magmatic processes differ between slow-spreading and fast-spreading crust. We present a new synthesis of ocean crust velocity structure, compiling observations from two-dimensional studies in the Atlantic, Pacific, and Indian ocean basins. The Layer 2/3 boundary was picked from each publication at a change in gradient either on velocity-depth functions or contour plots (with at least 0.5 km/s contour interval), or from the appropriate layer boundary for layered models. We picked multiple locations at each seismic refraction profile if warranted by model variability. Preliminary results show statistically significant differences in average Layer 2 and Layer 3 thicknesses between slow-spreading and superfast-spreading crust, with Layer 2 thinner and Layer 3 thicker for the higher spreading rate crust. The thickness changes are about equivalent, resulting in no change in mean crustal thickness. The Layer 2/3 boundary is often interpreted as the top of the gabbros; however, a comparison with mapped magma lens depths at the ridge axis shows that the boundary is typically deeper than average axial melt lens depth at superfast-spreading crust, and shallower at intermediate-spreading crust.

  4. Morning Martian Atmospheric Temperature Gradients and Fluctuations Observed by Mars Pathfinder

    NASA Technical Reports Server (NTRS)

    Mihalov, John D.; Haberle, R. M.; Murphy, J. R.; Seiff, A.; Wilson, G. R.

    1999-01-01

    We have studied the most prominent atmospheric temperature fluctuations observed during Martian mornings by Mars Pathfinder and have concluded, based on comparisons with wind directions, that they appear to be a result of atmospheric heating associated with the Lander spacecraft. Also, we have examined the morning surface layer temperature lapse rates, which are found to decrease as autumn approaches at the Pathfinder location, and which have mean (and median) values as large as 7.3 K/m in the earlier portions of the Pathfinder landed mission. It is plausible that brief isolated periods with gradients twice as steep are associated with atmospheric heating adjacent to Lander air bag material. In addition, we have calculated the gradient with height of the structure function obtained with Mars Pathfinder, for Mars' atmospheric temperatures measured within about 1.3 m from the surface, assuming a power law dependence, and have found that these gradients superficially resemble those reported for the upper region of the terrestrial stable boundary layer.

  5. Annular beam with segmented phase gradients

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Cheng, Shubo; Wu, Liang; Tao, Shaohua, E-mail: eshtao@csu.edu.cn

    2016-08-15

    An annular beam with a single uniform-intensity ring and multiple segments of phase gradients is proposed in this paper. Different from the conventional superposed vortices, such as the modulated optical vortices and the collinear superposition of multiple orbital angular momentum modes, the designed annular beam has a doughnut intensity distribution whose radius is independent of the phase distribution of the beam in the imaging plane. The phase distribution along the circumference of the doughnut beam can be segmented with different phase gradients. Similar to a vortex beam, the annular beam can also exert torques and rotate a trapped particle owingmore » to the orbital angular momentum of the beam. As the beam possesses different phase gradients, the rotation velocity of the trapped particle can be varied along the circumference. The simulation and experimental results show that an annular beam with three segments of different phase gradients can rotate particles with controlled velocities. The beam has potential applications in optical trapping and optical information processing.« less

  6. Combined solvent- and non-uniform temperature-programmed gradient liquid chromatography. I - A theoretical investigation.

    PubMed

    Gritti, Fabrice

    2016-11-18

    An new class of gradient liquid chromatography (GLC) is proposed and its performance is analyzed from a theoretical viewpoint. During the course of such gradients, both the solvent strength and the column temperature are simultaneously changed in time and space. The solvent and temperature gradients propagate along the chromatographic column at their own and independent linear velocity. This class of gradient is called combined solvent- and temperature-programmed gradient liquid chromatography (CST-GLC). The general expressions of the retention time, retention factor, and of the temporal peak width of the analytes at elution in CST-GLC are derived for linear solvent strength (LSS) retention models, modified van't Hoff retention behavior, linear and non-distorted solvent gradients, and for linear temperature gradients. In these conditions, the theory predicts that CST-GLC is equivalent to a unique and apparent dynamic solvent gradient. The apparent solvent gradient steepness is the sum of the solvent and temperature steepness. The apparent solvent linear velocity is the reciprocal of the steepness-averaged sum of the reciprocal of the actual solvent and temperature linear velocities. The advantage of CST-GLC over conventional GLC is demonstrated for the resolution of protein digests (peptide mapping) when applying smooth, retained, and linear acetonitrile gradients in combination with a linear temperature gradient (from 20°C to 90°C) using 300μm×150mm capillary columns packed with sub-2 μm particles. The benefit of CST-GLC is demonstrated when the temperature gradient propagates at the same velocity as the chromatographic speed. The experimental proof-of-concept for the realization of temperature ramps propagating at a finite and constant linear velocity is also briefly described. Copyright © 2016 Elsevier B.V. All rights reserved.

  7. Gradients estimation from random points with volumetric tensor in turbulence

    NASA Astrophysics Data System (ADS)

    Watanabe, Tomoaki; Nagata, Koji

    2017-12-01

    We present an estimation method of fully-resolved/coarse-grained gradients from randomly distributed points in turbulence. The method is based on a linear approximation of spatial gradients expressed with the volumetric tensor, which is a 3 × 3 matrix determined by a geometric distribution of the points. The coarse grained gradient can be considered as a low pass filtered gradient, whose cutoff is estimated with the eigenvalues of the volumetric tensor. The present method, the volumetric tensor approximation, is tested for velocity and passive scalar gradients in incompressible planar jet and mixing layer. Comparison with a finite difference approximation on a Cartesian grid shows that the volumetric tensor approximation computes the coarse grained gradients fairly well at a moderate computational cost under various conditions of spatial distributions of points. We also show that imposing the solenoidal condition improves the accuracy of the present method for solenoidal vectors, such as a velocity vector in incompressible flows, especially when the number of the points is not large. The volumetric tensor approximation with 4 points poorly estimates the gradient because of anisotropic distribution of the points. Increasing the number of points from 4 significantly improves the accuracy. Although the coarse grained gradient changes with the cutoff length, the volumetric tensor approximation yields the coarse grained gradient whose magnitude is close to the one obtained by the finite difference. We also show that the velocity gradient estimated with the present method well captures the turbulence characteristics such as local flow topology, amplification of enstrophy and strain, and energy transfer across scales.

  8. Metallicity gradient of the thick disc progenitor at high redshift

    NASA Astrophysics Data System (ADS)

    Kawata, Daisuke; Allende Prieto, Carlos; Brook, Chris B.; Casagrande, Luca; Ciucă, Ioana; Gibson, Brad K.; Grand, Robert J. J.; Hayden, Michael R.; Hunt, Jason A. S.

    2018-01-01

    We have developed a novel Markov Chain Monte Carlo chemical 'painting' technique to explore possible radial and vertical metallicity gradients for the thick disc progenitor. In our analysis, we match an N-body simulation to the data from the Apache Point Observatory Galactic Evolution Experiment survey. We assume that the thick disc has a constant scaleheight and has completed its formation at an early epoch, after which time radial mixing of its stars has taken place. Under these assumptions, we find that the initial radial metallicity gradient of the thick disc progenitor should not be negative, but either flat or even positive, to explain the current negative vertical metallicity gradient of the thick disc. Our study suggests that the thick disc was built-up in an inside-out and upside-down fashion, and older, smaller and thicker populations are more metal poor. In this case, star-forming discs at different epochs of the thick disc formation are allowed to have different radial metallicity gradients, including a negative one, which helps to explain a variety of slopes observed in high-redshift disc galaxies. This scenario helps to explain the positive slope of the metallicity-rotation velocity relation observed for the Galactic thick disc. On the other hand, radial mixing flattens the slope of an existing gradient.

  9. Image Motion Detection And Estimation: The Modified Spatio-Temporal Gradient Scheme

    NASA Astrophysics Data System (ADS)

    Hsin, Cheng-Ho; Inigo, Rafael M.

    1990-03-01

    The detection and estimation of motion are generally involved in computing a velocity field of time-varying images. A completely new modified spatio-temporal gradient scheme to determine motion is proposed. This is derived by using gradient methods and properties of biological vision. A set of general constraints is proposed to derive motion constraint equations. The constraints are that the second directional derivatives of image intensity at an edge point in the smoothed image will be constant at times t and t+L . This scheme basically has two stages: spatio-temporal filtering, and velocity estimation. Initially, image sequences are processed by a set of oriented spatio-temporal filters which are designed using a Gaussian derivative model. The velocity is then estimated for these filtered image sequences based on the gradient approach. From a computational stand point, this scheme offers at least three advantages over current methods. The greatest advantage of the modified spatio-temporal gradient scheme over the traditional ones is that an infinite number of motion constraint equations are derived instead of only one. Therefore, it solves the aperture problem without requiring any additional assumptions and is simply a local process. The second advantage is that because of the spatio-temporal filtering, the direct computation of image gradients (discrete derivatives) is avoided. Therefore the error in gradients measurement is reduced significantly. The third advantage is that during the processing of motion detection and estimation algorithm, image features (edges) are produced concurrently with motion information. The reliable range of detected velocity is determined by parameters of the oriented spatio-temporal filters. Knowing the velocity sensitivity of a single motion detection channel, a multiple-channel mechanism for estimating image velocity, seldom addressed by other motion schemes in machine vision, can be constructed by appropriately choosing and

  10. Detection of Intracluster Gas Bulk Velocities in the Perseus and Centaurus Clusters

    NASA Astrophysics Data System (ADS)

    Dupke, Renato A.; Bregman, Joel N.

    We report the results of spatially resolved X-ray spectroscopy of 8 different ASCApointings distributed symmetrically around the center of the Perseus cluster. The outer region of the intracluster gas is roughly isothermal, with temperature ~ 6-7 keV, and metal abundance ~ 0.3 Solar. Spectral analysis of the central pointing is consistent with the presence of a cooling flow and a central metal abundance gradient. A significant velocity gradient is found along an axis highly discrepant with the major axis of the X-ray elongation. The radial velocity difference is found to be greater than 1000 km s-1Mpc-1 at the 90% confidence level. Simultaneous fittings of GIS 2 & 3 indicate that two symmetrically opposed regions have different radial velocities at the 95% confidence level and the F-test rules out constant velocities for these regions at the 99% level. Intrinsic short and long term variations of gain are unlikely (P < 0.03) to explain the velocity discrepancies. We also report the preliminary results of a similar analysis carried out for the Centaurus cluster, where long-exposure SIS data is available. We also find a significant velocity gradient near the central regions (3'-8' of Centaurus. If attributed to bulk rotation the correspondent circular velocity is ~1500±150 km s-1 (at 90% confidence). The line of maximum velocity gradient in Centaurus is near-perpendicular to the infalling galaxy group associated with NGC 4709.

  11. Whistler Waves Driven by Anisotropic Strahl Velocity Distributions: Cluster Observations

    NASA Technical Reports Server (NTRS)

    Vinas, A.F.; Gurgiolo, C.; Nieves-Chinchilla, T.; Gary, S. P.; Goldstein, M. L.

    2010-01-01

    Observed properties of the strahl using high resolution 3D electron velocity distribution data obtained from the Cluster/PEACE experiment are used to investigate its linear stability. An automated method to isolate the strahl is used to allow its moments to be computed independent of the solar wind core+halo. Results show that the strahl can have a high temperature anisotropy (T(perpindicular)/T(parallell) approximately > 2). This anisotropy is shown to be an important free energy source for the excitation of high frequency whistler waves. The analysis suggests that the resultant whistler waves are strong enough to regulate the electron velocity distributions in the solar wind through pitch-angle scattering

  12. Observation of Enhanced Hole Extraction in Br Concentration Gradient Perovskite Materials.

    PubMed

    Kim, Min-Cheol; Kim, Byeong Jo; Son, Dae-Yong; Park, Nam-Gyu; Jung, Hyun Suk; Choi, Mansoo

    2016-09-14

    Enhancing hole extraction inside the perovskite layer is the key factor for boosting photovoltaic performance. Realization of halide concentration gradient perovskite materials has been expected to exhibit rapid hole extraction due to the precise bandgap tuning. Moreover, a formation of Br-rich region on the tri-iodide perovskite layer is expected to enhance moisture stability without a loss of current density. However, conventional synthetic techniques of perovskite materials such as the solution process have not achieved the realization of halide concentration gradient perovskite materials. In this report, we demonstrate the fabrication of Br concentration gradient mixed halide perovskite materials using a novel and facile halide conversion method based on vaporized hydrobromic acid. Accelerated hole extraction and enhanced lifetime due to Br gradient was verified by observing photoluminescence properties. Through the combination of secondary ion mass spectroscopy and transmission electron microscopy with energy-dispersive X-ray spectroscopy analysis, the diffusion behavior of Br ions in perovskite materials was investigated. The Br-gradient was found to be eventually converted into a homogeneous mixed halide layer after undergoing an intermixing process. Br-substituted perovskite solar cells exhibited a power conversion efficiency of 18.94% due to an increase in open circuit voltage from 1.08 to 1.11 V and an advance in fill-factor from 0.71 to 0.74. Long-term stability was also dramatically enhanced after the conversion process, i.e., the power conversion efficiency of the post-treated device has remained over 97% of the initial value under high humid conditions (40-90%) without any encapsulation for 4 weeks.

  13. Simulations of star-forming molecular clouds: observational predictions

    NASA Astrophysics Data System (ADS)

    Zhang, Shangjia; Hartmann, Lee; Kuznetsova, Aleksandra; Abelardo Zamora, Manuel

    2018-01-01

    Observations of protostellar molecular cloud cores can be used to test theories of star formation. However, observational results can be biased because of limited information: (a) only two spatial dimensions and one velocity dimension can be measured, (b) and cores generally are not spherically symmetric. We use numerical simulations of the formation and collapse of molecular gas with sink particles to make observational predictions. We use the radiative transfer code LIME to predict CO and NH3 channel maps. We find reasonable agreement with observed velocity structures and gradients but occasional large differences depending on viewing angle.

  14. Influence of Flow Gradients on Mach Stem Initiation of PBX-9502

    NASA Astrophysics Data System (ADS)

    Hull, Lawrence; Miller, Phillip; Mas, Eric; Focused Experiments Team

    2017-06-01

    Recent experiments and theory explore the effect of flow gradients on reaction acceleration and stability in the pressure-enhanced region between colliding sub-detonative shock waves in PBX-9502. The experiments are designed to produce divergent curved incident shock waves that interact in a convergent irregular reflection, or ``Mach stem'', configuration. Although this flow is fundamentally unsteady, such a configuration does feature particle paths having a single shock wave that increases the pressure from zero to the wave-reflected enhanced pressure. Thus, the possibility of pre-shock desensitization is precluded in this interaction region. Diagnostics record arrival wave velocity, shape, and material velocity along the angled free surface face of a large wedge. The wedge is large enough to allow observation of the wave structure for distances much larger than the run-to-detonation derived from classical ``Pop plot'' data. The explosive driver system produces the incident shocks and allows some control of the flow gradients in the collision region. Further, the incident shocks are very weak and do not transition to detonation. The experiments discussed feature incident shock waves that would be expected to cause initiation in the Mach stem, based on the Pop plot. Results show that the introduction of pressure/velocity gradients in the reaction zone strongly influences the ability of the flow to build to a steady ``CJ'' detonation. As expected, the ability of the Mach stem to stabilize or accelerate is strongly influenced by the incident shock pressure.

  15. How can we Optimize Global Satellite Observations of Glacier Velocity and Elevation Changes?

    NASA Astrophysics Data System (ADS)

    Willis, M. J.; Pritchard, M. E.; Zheng, W.

    2015-12-01

    We have started a global compilation of glacier surface elevation change rates measured by altimeters and differencing of Digital Elevation Models and glacier velocities measured by Synthetic Aperture Radar (SAR) and optical feature tracking as well as from Interferometric SAR (InSAR). Our goal is to compile statistics on recent ice flow velocities and surface elevation change rates near the fronts of all available glaciers using literature and our own data sets of the Russian Arctic, Patagonia, Alaska, Greenland and Antarctica, the Himalayas, and other locations. We quantify the percentage of the glaciers on the planet that can be regarded as fast flowing glaciers, with surface velocities of more than 50 meters per year, while also recording glaciers that have elevation change rates of more than 2 meters per year. We examine whether glaciers have significant interannual variations in velocities, or have accelerated or stagnated where time series of ice motions are available. We use glacier boundaries and identifiers from the Randolph Glacier Inventory. Our survey highlights glaciers that are likely to react quickly to changes in their mass accumulation rates. The study also identifies geographical areas where our knowledge of glacier dynamics remains poor. Our survey helps guide how frequently observations must be made in order to provide quality satellite-derived velocity and ice elevation observations at a variety of glacier thermal regimes, speeds and widths. Our objectives are to determine to what extent the joint NASA and Indian Space Research Organization Synthetic Aperture Radar mission (NISAR) will be able to provide global precision coverage of ice speed changes and to determine how to optimize observations from the global constellation of satellite missions to record important changes to glacier elevations and velocities worldwide.

  16. Observations of wave-induced pore pressure gradients and bed level response on a surf zone sandbar

    NASA Astrophysics Data System (ADS)

    Anderson, Dylan; Cox, Dan; Mieras, Ryan; Puleo, Jack A.; Hsu, Tian-Jian

    2017-06-01

    Horizontal and vertical pressure gradients may be important physical mechanisms contributing to onshore sediment transport beneath steep, near-breaking waves in the surf zone. A barred beach was constructed in a large-scale laboratory wave flume with a fixed profile containing a mobile sediment layer on the crest of the sandbar. Horizontal and vertical pore pressure gradients were obtained by finite differences of measurements from an array of pressure transducers buried within the upper several centimeters of the bed. Colocated observations of erosion depth were made during asymmetric wave trials with wave heights between 0.10 and 0.98 m, consistently resulting in onshore sheet flow sediment transport. The pore pressure gradient vector within the bed exhibited temporal rotations during each wave cycle, directed predominantly upward under the trough and then rapidly rotating onshore and downward as the wavefront passed. The magnitude of the pore pressure gradient during each phase of rotation was correlated with local wave steepness and relative depth. Momentary bed failures as deep as 20 grain diameters were coincident with sharp increases in the onshore-directed pore pressure gradients, but occurred at horizontal pressure gradients less than theoretical critical values for initiation of the motion for compact beds. An expression combining the effects of both horizontal and vertical pore pressure gradients with bed shear stress and soil stability is used to determine that failure of the bed is initiated at nonnegligible values of both forces.Plain Language SummaryThe pressure <span class="hlt">gradient</span> present within the seabed beneath breaking waves may be an important physical mechanism transporting sediment. A large-scale laboratory was used to replicate realistic surfzone conditions in controlled tests, allowing for horizontal and vertical pressure <span class="hlt">gradient</span> magnitudes and the resulting sediment bed response to be <span class="hlt">observed</span> with</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19810038514&hterms=1587&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231587','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810038514&hterms=1587&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231587"><span>Latitude dependence of solar wind <span class="hlt">velocity</span> <span class="hlt">observed</span> at not less than 1 AU</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mitchell, D. G.; Roelof, E. C.; Wolfe, J. H.</p> <p>1981-01-01</p> <p>The large-scale solar wind <span class="hlt">velocity</span> structure in the outer heliosphere has been systematically analyzed for Carrington rotations 1587-1541 (March 1972 to April 1976). Spacecraft data were taken from Imp 7/8 at earth, Pioneer 6, 8, and 9 near 1 AU, and Pioneer 10 and 11 between 1.6 and 5 AU. Using the constant radial <span class="hlt">velocity</span> solar wind approximation to map all of the <span class="hlt">velocity</span> data to its high coronal emission heliolongitude, the <span class="hlt">velocity</span> structure <span class="hlt">observed</span> at different spacecraft was examined for latitudinal dependence and compared with coronal structure in soft X-rays and H-alpha absorption features. The constant radial <span class="hlt">velocity</span> approximation usually remains self-consistent in decreasing or constant <span class="hlt">velocity</span> solar wind out to 5 AU, enabling us to separate radial from latitudinal propagation effects. Several examples of sharp nonmeridional stream boundaries in interplanetary space (about 5 deg latitude in width), often directly associated with features in coronal X-rays and H-alpha were found.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70023518','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70023518"><span>A simple algorithm for sequentially incorporating gravity <span class="hlt">observations</span> in seismic traveltime tomography</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Parsons, T.; Blakely, R.J.; Brocher, T.M.</p> <p>2001-01-01</p> <p>The geologic structure of the Earth's upper crust can be revealed by modeling variation in seismic arrival times and in potential field measurements. We demonstrate a simple method for sequentially satisfying seismic traveltime and <span class="hlt">observed</span> gravity residuals in an iterative 3-D inversion. The algorithm is portable to any seismic analysis method that uses a gridded representation of <span class="hlt">velocity</span> structure. Our technique calculates the gravity anomaly resulting from a <span class="hlt">velocity</span> model by converting to density with Gardner's rule. The residual between calculated and <span class="hlt">observed</span> gravity is minimized by weighted adjustments to the model <span class="hlt">velocity</span>-depth <span class="hlt">gradient</span> where the <span class="hlt">gradient</span> is steepest and where seismic coverage is least. The adjustments are scaled by the sign and magnitude of the gravity residuals, and a smoothing step is performed to minimize vertical streaking. The adjusted model is then used as a starting model in the next seismic traveltime iteration. The process is repeated until one <span class="hlt">velocity</span> model can simultaneously satisfy both the gravity anomaly and seismic traveltime <span class="hlt">observations</span> within acceptable misfits. We test our algorithm with data gathered in the Puget Lowland of Washington state, USA (Seismic Hazards Investigation in Puget Sound [SHIPS] experiment). We perform resolution tests with synthetic traveltime and gravity <span class="hlt">observations</span> calculated with a checkerboard <span class="hlt">velocity</span> model using the SHIPS experiment geometry, and show that the addition of gravity significantly enhances resolution. We calculate a new <span class="hlt">velocity</span> model for the region using SHIPS traveltimes and <span class="hlt">observed</span> gravity, and show examples where correlation between surface geology and modeled subsurface <span class="hlt">velocity</span> structure is enhanced.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25e6312R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25e6312R"><span>Measurements of ion <span class="hlt">velocity</span> separation and ionization in multi-species plasma shocks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rinderknecht, Hans G.; Park, H.-S.; Ross, J. S.; Amendt, P. A.; Wilks, S. C.; Katz, J.; Hoffman, N. M.; Kagan, G.; Vold, E. L.; Keenan, B. D.; Simakov, A. N.; Chacón, L.</p> <p>2018-05-01</p> <p>The ion <span class="hlt">velocity</span> structure of a strong collisional shock front in a plasma with multiple ion species is directly probed in laser-driven shock-tube experiments. Thomson scattering of a 263.25 nm probe beam is used to diagnose ion composition, temperature, and flow <span class="hlt">velocity</span> in strong shocks ( M ˜6 ) propagating through low-density ( ρ˜0.1 mg/cc) plasmas composed of mixtures of hydrogen (98%) and neon (2%). Within the preheat region of the shock front, two <span class="hlt">velocity</span> populations of ions are <span class="hlt">observed</span>, a characteristic feature of strong plasma shocks. The ionization state of the Ne is <span class="hlt">observed</span> to change within the shock front, demonstrating an ionization-timescale effect on the shock front structure. The forward-streaming proton feature is shown to be unexpectedly cool compared to predictions from ion Fokker-Planck simulations; the neon ionization <span class="hlt">gradient</span> is evaluated as a possible cause.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E.244B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E.244B"><span>Matching Lithosphere <span class="hlt">velocity</span> changes to the GOCE gravity signal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Braitenberg, Carla</p> <p>2016-07-01</p> <p>Authors: Carla Braitenberg, Patrizia Mariani, Alberto Pastorutti Department of Mathematics and Geosciences, University of Trieste Via Weiss 1, 34100 Trieste Seismic tomography models result in 3D <span class="hlt">velocity</span> models of lithosphere and sublithospheric mantle, which are due to mineralogic compositional changes and variations in the thermal <span class="hlt">gradient</span>. The assignment of density is non-univocal and can lead to inverted density changes with respect to <span class="hlt">velocity</span> changes, depending on composition and temperature. <span class="hlt">Velocity</span> changes due to temperature result in a proportional density change, whereas changes due to compositional changes and age of the lithosphere can lead to density changes of inverted sign. The relation between <span class="hlt">velocity</span> and density implies changes in the lithosphere rigidity. We analyze the GOCE <span class="hlt">gradient</span> fields and the <span class="hlt">velocity</span> models jointly, making simulations on thermal and compositional density changes, using the <span class="hlt">velocity</span> models as constraint on lithosphere geometry. The correlations are enhanced by applying geodynamic plate reconstructions to the GOCE gravity field and the tomography models which places today's <span class="hlt">observed</span> fields at the Gondwana pre-breakup position. We find that the lithosphere geometry is a controlling factor on the overlying geologic elements, defining the regions where rifting and collision alternate and repeat through time. The study is carried out globally, with focus on the conjugate margins of the African and South American continents. The background for the study can be found in the following publications where the techniques which have been used are described: Braitenberg, C., Mariani, P. and De Min, A. (2013). The European Alps and nearby orogenic belts sensed by GOCE, Boll. Bollettino di Geofisica Teorica ed Applicata, 54(4), 321-334. doi:10.4430/bgta0105---- Braitenberg, C. and Mariani, P. (2015). Geological implications from complete Gondwana GOCE-products reconstructions and link to lithospheric roots. Proceedings of 5th</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.C23C0669E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.C23C0669E"><span>Satellite <span class="hlt">Observations</span> of Glacier Surface <span class="hlt">Velocities</span> in Southeast Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Elliott, J.; Melkonian, A. K.; Pritchard, M. E.</p> <p>2012-12-01</p> <p>Glaciers in southeast Alaska are undergoing rapid changes and are significant contributors to sea level rise. A key to understanding the ice dynamics is knowledge of the surface <span class="hlt">velocities</span>, which can be used with ice thickness measurements to derive mass flux rates. For many glaciers in Alaska, surface <span class="hlt">velocity</span> estimates either do not exist or are based on data that are at least a decade old. Here we present updated maps of glacier surface <span class="hlt">velocities</span> in southeast Alaska produced through a pixel tracking technique using synthetic aperture radar data and high-resolution optical imagery. For glaciers with previous <span class="hlt">velocity</span> estimates, we will compare the results and discuss possible implications for ice dynamics. We focus on Glacier Bay and the Stikine Icefield, which contain a number of fast-flowing tidewater glaciers including LeConte, Johns Hopkins, and La Perouse. For the Johns Hopkins, we will also examine the influence a massive landslide in June 2012 had on flow dynamics. Our <span class="hlt">velocity</span> maps show that within Glacier Bay, the highest surface <span class="hlt">velocities</span> occur on the tidewater glaciers. La Perouse, the only Glacier Bay glacier to calve directly into the Pacific Ocean, has maximum <span class="hlt">velocities</span> of 3.5 - 4 m/day. Johns Hopkins Glacier shows 4 m/day <span class="hlt">velocities</span> at both its terminus and in its upper reaches, with lower <span class="hlt">velocities</span> of ~1-3 m/day in between those two regions. Further north, the Margerie Glacier has a maximum <span class="hlt">velocity</span> of ~ 4.5 m/day in its upper reaches and a <span class="hlt">velocity</span> of ~ 2 m/day at its terminus. Along the Grand Pacific terminus, the western terminus fed by the Ferris Glacier displays <span class="hlt">velocities</span> of about 1 m/day while the eastern terminus has lower <span class="hlt">velocities</span> of < 0.5 m/day. The lake terminating glaciers along the Pacific coast have overall lower surface <span class="hlt">velocities</span>, but they display complex flow patterns. The Alsek Glacier displays maximum <span class="hlt">velocities</span> of 2.5 m/day above where it divides into two branches. <span class="hlt">Velocities</span> at the terminus of the northern branch reach 1</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T41C2940C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T41C2940C"><span>Salt Interval <span class="hlt">Velocities</span> vs Latitude in the Deepwater Gulf of Mexico: Keathley Canyon and Walker Ridge Areas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cornelius, S.; Castagna, J. P.</p> <p>2016-12-01</p> <p>ABSTRACT A well log database of approximately 300 well logs from the Keathley Canyon and Walker Ridge areas of the Gulf of Mexico plus Mad Dog Field and Mission Deep Field in Green Canyon has been created for the purpose of building a geologically based 3D <span class="hlt">velocity</span> model. While in the process of calibrating the finished <span class="hlt">velocity</span> model, a scatter plot was made of all salt interval <span class="hlt">velocities</span> versus latitude and an unexpected correlation was <span class="hlt">observed</span>. Five different interval <span class="hlt">velocity</span> zones have been identified with each having certain associated mineralogies within a latitude range. The salt interval <span class="hlt">velocity</span> in the southern limits of the study area is higher than 15,000 ft/sec (4572 m/sec) due to the presence of gypsum. The northern most wells in the project area have anhydrite present inside the salt matrix such that their interval <span class="hlt">velocity</span> can be as high as 18,535 ft/sec (5650 m/sec). In the mid-latitude zones, sylvite, siltstone, claystone, shale, tar and bitumen, with small traces of both anhydrite and gypsum, are found within the salt, yielding salt interval <span class="hlt">velocity</span> variation from 14,388 ft/sec to 14,909 ft/sec (4386 m/sec to 4544 m/sec). The mineralogical content of the salt in each well was roughly estimated from mud logs and the corresponding interval <span class="hlt">velocities</span> were determined from vertical seismic profiles, checkshot surveys, and sonic logs. Both geothermal <span class="hlt">gradients</span> and overburden geopressure <span class="hlt">gradients</span> between the mudline and the true vertical depth at well bottom calculated from this well database do not show the same correlation with latitude as the salt interval <span class="hlt">velocities</span>. Mineralogical modeling of the salt composition using Hashin-Shtrikman bounds shows that these various inclusions within the salt matrix can be the cause of the <span class="hlt">observed</span> variations in the salt interval <span class="hlt">velocities</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.S43B0999T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.S43B0999T"><span>Acoustic <span class="hlt">Velocity</span> Of The Sediments Offshore Southwestern Taiwan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsai, C.; Liu, C.; Huang, P.</p> <p>2004-12-01</p> <p>Along the Manila Trench south of 21øXN, deep-sea sediments are being underthrusted beneath the Taiwan accretionary prism which is composed of the Kaoping Slope and Hengchun Ridge. Offshore southwestern Taiwan, foreland sediments and Late Miocene strata of the Tainan Basin are being accreted onto the fold-and thrust belt of the syn-collision accretionary wedge of the Kaoping Slope. The Kaoping Slope consists of thick Neogene to Recent siliciclastics deformed by fold-and-thrust structures and mud diapers. These Pliocene-Quaternary sediments deposited in the Kaoping Shelf and upper slope area are considered to be paleo-channel deposits confined by NNE-SSW trend mud diapiric structure. Seismic P-wave <span class="hlt">velocities</span> of the sediment deposited in the Kaoping Shelf and Kaoping Slope area are derived from mutichannel seismic reflection data and wide-angle reflection and refraction profiles collected by sonobuoys. Sediment <span class="hlt">velocity</span> structures constrained from mutichannel seismic reflection data using <span class="hlt">velocity</span> spectrum analysis method and that derived from sonobuoy data using tau-sum inversion method are compared, and they both provide consistent <span class="hlt">velocity</span> structures. Seismic <span class="hlt">velocities</span> were analyzed along the seismic profile from the surface to maximum depths of about 2.0 km below the seafloor. Our model features a sediment layer1 with 400 ms in thickness and a sediment layer2 with 600 ms in thickness. For the shelf sediments, we <span class="hlt">observe</span> a linear interval <span class="hlt">velocity</span> trend of V=1.53+1.91T in layer1, and V=1.86+0.87T in layer2, where T is the one way travel time within the layer. For the slop sediment, the trend of V=1.47+1.93T in layer1, and V=1.70+1.55T in layer2. The layer1¡¦s <span class="hlt">velocities</span> <span class="hlt">gradients</span> are similar between the shelf (1.91 km/sec2) and the slope(1.93 km/sec2). It means layer1 distributes over the slope and shelf widely. The result of the sediment <span class="hlt">velocity</span> <span class="hlt">gradients</span> in this area are in good agreement with that reported for the south Atlantic continental margins.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.1379F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.1379F"><span>Limitations on Inferring 3D Architecture and Dynamics From Surface <span class="hlt">Velocities</span> in the India-Eurasia Collision Zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Flesch, L.; Bendick, R.; Bischoff, S.</p> <p>2018-02-01</p> <p>Surface <span class="hlt">velocities</span> derived from Global Positioning System <span class="hlt">observations</span> and Quaternary fault slip rates measured throughout an extended region of high topography in South Asia vary smoothly over thousands of kilometers and are broadly symmetrical, with components of both north-south shortening and east-west extension relative to stable Eurasia. The <span class="hlt">observed</span> <span class="hlt">velocity</span> field does not contain discontinuities or steep <span class="hlt">gradients</span> attributable to along-strike differences in collision architecture, despite the well-documented presence of a lithospheric slab beneath the Pamir but not the Tibetan Plateau. We use a modified Akaike information criterion (AICc) to show that surface <span class="hlt">velocities</span> do not efficiently constrain 3D rheology, geometry, or force balance. Therefore, although other geophysical and geological <span class="hlt">observations</span> may indicate the presence of mechanical or dynamic heterogeneities within the Indian-Asian collision, the surface Global Positioning System <span class="hlt">velocities</span> contain little or no usable information about them.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1953n0139K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1953n0139K"><span>Study of electrostatic electron cyclotron parallel flow <span class="hlt">velocity</span> shear instability in the magnetosphere of Saturn</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kandpal, Praveen; Pandey, R. S.</p> <p>2018-05-01</p> <p>In the present paper, the study of electrostatic electron cyclotron parallel flow <span class="hlt">velocity</span> shear instability in presence of perpendicular inhomogeneous DC electric field has been carried out in the magnetosphere of Saturn. Dimensionless growth rate variation of electron cyclotron waves has been <span class="hlt">observed</span> with respect to k⊥ ρe for various plasma parameters. Effect of <span class="hlt">velocity</span> shear scale length (Ae), inhomogeneity (P/a), the ratio of ion to electron temperature (Ti/Te) and density <span class="hlt">gradient</span> (ɛnρe) on the growth of electron cyclotron waves in the inner magnetosphere of Saturn has been studied and analyzed. The mathematical formulation and computation of dispersion relation and growth rate have been done by using the method of characteristic solution and kinetic approach. This theoretical analysis has been done taking the relevant data from the Cassini spacecraft in the inner magnetosphere of Saturn. We have considered ambient magnetic field data and other relevant data for this study at the radial distance of ˜4.82-5.00 Rs. In our study <span class="hlt">velocity</span> shear and ion to electron temperature ratio have been <span class="hlt">observed</span> to be the major sources of free energy for the electron cyclotron instability. The inhomogeneity of electric field caused a small noticeable impact on the growth rate of electrostatic electron cyclotron instability. Density <span class="hlt">gradient</span> has been <span class="hlt">observed</span> playing stabilizing effect on electron cyclotron instability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AdSpR..61..581K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AdSpR..61..581K"><span><span class="hlt">Velocity</span> shear Kelvin-Helmholtz instability with inhomogeneous DC electric field in the magnetosphere of Saturn</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kandpal, Praveen; Kaur, Rajbir; Pandey, R. S.</p> <p>2018-01-01</p> <p>In this paper parallel flow <span class="hlt">velocity</span> shear Kelvin-Helmholtz instability has been studied in two different extended regions of the inner magnetosphere of Saturn. The method of the characteristic solution and kinetic approach has been used in the mathematical calculation of dispersion relation and growth rate of K-H waves. Effect of magnetic field (B), inhomogeneity (P/a), <span class="hlt">velocity</span> shear scale length (Ai), temperature anisotropy (T⊥ /T||), electric field (E), ratio of electron to ion temperature (Te /Ti), density <span class="hlt">gradient</span> (εnρi) and angle of propagation (θ) on the dimensionless growth rate of K-H waves in the inner magnetosphere of Saturn has been <span class="hlt">observed</span> with respect to k⊥ρi . Calculations of this theoretical analysis have been done taking the data from the Cassini in the inner magnetosphere of Saturn in the two extended regions of Rs ∼4.60-4.01 and Rs ∼4.82-5.0. In our study <span class="hlt">velocity</span> shear, temperature anisotropy and magnitude of the electric field are <span class="hlt">observed</span> to be the major sources of free energy for the K-H instability in both the regions considered. The inhomogeneity of electric field, electron-ion temperature ratio, and density <span class="hlt">gradient</span> have been <span class="hlt">observed</span> playing stabilizing effect on K-H instability. This study also indicates the effect of the vicinity of icy moon Enceladus on the growth of K-H instability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016DPS....4812115J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016DPS....4812115J"><span>Collision of Dual Aggregates (CODA): Experimental <span class="hlt">observations</span> of low-<span class="hlt">velocity</span> collisions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jorges, Jeffery; Dove, Adrienne; Colwell, Josh E.</p> <p>2016-10-01</p> <p>Low-<span class="hlt">velocity</span> collisions are one of the driving factors that determine the particle size distribution and particle size evolution in planetary ring systems and in the early stages of planet formation. Collisions of sub-micron to decimeter-sized objects may result in particle growth by accretion, rebounding, or erosive processes that result in the production of additional smaller particles. Numerical simulations of these systems are limited by a need to understand these collisional parameters over a range of conditions. We present the results of a sequence of laboratory experiments designed to explore collisions over a range of parameter space . We are able to <span class="hlt">observe</span> low-<span class="hlt">velocity</span> collisions by conducting experiments in vacuum chambers in our 0.8-sec drop tower apparatus. Initial experiments utilize a variety of impacting spheres, including glass, Teflon, aluminum, stainless steel, and brass. These spheres are either used in their natural state or are "mantled" - coated with a few-mm thick layer of a cohesive powder. A high-speed, high-resolution video camera is used to record the motion of the colliding bodies. We track the particles to determine impactor speeds before and after collision, the impact parameter, and the collisional outcome. In the case of the mantled impactors, we can assess how much rotation is generated by the collision and estimate how much powder is released (i.e. how much mass is lost) due to the collision. We also determine how the coefficient of restitution varies as a function of material type, morphology, and impact <span class="hlt">velocity</span>. With impact <span class="hlt">velocities</span> ranging from about 20-100 cm/s we <span class="hlt">observe</span> that mantling of particles significantly reduces their coefficients of restitution, but we see basically no dependence of the coefficient of restitution on the impact <span class="hlt">velocity</span>, impact parameter, or system mass. The results of this study will contribute to a better empirical model of collisional outcomes that will be refined with numerical simulation of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ASPC..458..185R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ASPC..458..185R"><span>The Galactic Bulge Radial <span class="hlt">Velocity</span>/Abundance Assay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rich, R. M.</p> <p>2012-08-01</p> <p>The Bulge Radial <span class="hlt">Velocity</span> Assay (BRAVA) measured radial <span class="hlt">velocities</span> for ˜ 9500 late-type giants in the Galactic bulge, predominantly from -10° < l < +10° and -2° < b < -10°. The project has discovered that the bulge exhibits cylindrical rotation characteristic of bars, and two studies of dynamics (Shen et al. 2010; Wang et al. 2012 MNRAS sub.) find that bar models- either N-body formed from an instability in a preexisting disk, or a self-consistent model- can account for the <span class="hlt">observed</span> kinematics. Studies of the Plaut field at (l,b) = 0°, -8° show that alpha enhancement is found in bulge giants even 1 kpc from the nucleus. New infrared studies extending to within 0.25° = 35 pc of the Galactic Center find no iron or alpha <span class="hlt">gradient</span> from Baade's Window (l,b) = 0.9°, -3.9° to our innermost field, in contrast to the marked <span class="hlt">gradient</span> <span class="hlt">observed</span> in the outer bulge. We consider the case of the remarkable globular cluster Terzan 5, which has a strongly bimodal iron and rm [α/Fe] within its members, and we consider evidence pro and con that the bulge was assembled from dissolved clusters. The Subaru telescope has the potential to contribute to study of the Galactic bulge, especially using the Hyper Superime-Cam and planned spectroscopic modes, as well as the high resolution spectrograph. The planned Jasmine satellite series may deliver a comprehensive survey of distances and proper motions of bulge stars, and insight into the origin and importance of the X-shaped bulge.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1344234-convective-cloud-vertical-velocity-mass-flux-characteristics-from-radar-wind-profiler-observations-during-goamazon2014-vertical-velocity-goamazon2014','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1344234-convective-cloud-vertical-velocity-mass-flux-characteristics-from-radar-wind-profiler-observations-during-goamazon2014-vertical-velocity-goamazon2014"><span>Convective cloud vertical <span class="hlt">velocity</span> and mass-flux characteristics from radar wind profiler <span class="hlt">observations</span> during GoAmazon2014/5: VERTICAL <span class="hlt">VELOCITY</span> GOAMAZON2014/5</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Giangrande, Scott E.; Toto, Tami; Jensen, Michael P.; ...</p> <p>2016-11-15</p> <p>A radar wind profiler data set collected during the 2 year Department of Energy Atmospheric Radiation Measurement <span class="hlt">Observations</span> and Modeling of the Green Ocean Amazon (GoAmazon2014/5) campaign is used to estimate convective cloud vertical <span class="hlt">velocity</span>, area fraction, and mass flux profiles. Vertical <span class="hlt">velocity</span> <span class="hlt">observations</span> are presented using cumulative frequency histograms and weighted mean profiles to provide insights in a manner suitable for global climate model scale comparisons (spatial domains from 20 km to 60 km). Convective profile sensitivity to changes in environmental conditions and seasonal regime controls is also considered. Aggregate and ensemble average vertical <span class="hlt">velocity</span>, convective area fraction, andmore » mass flux profiles, as well as magnitudes and relative profile behaviors, are found consistent with previous studies. Updrafts and downdrafts increase in magnitude with height to midlevels (6 to 10 km), with updraft area also increasing with height. Updraft mass flux profiles similarly increase with height, showing a peak in magnitude near 8 km. Downdrafts are <span class="hlt">observed</span> to be most frequent below the freezing level, with downdraft area monotonically decreasing with height. Updraft and downdraft profile behaviors are further stratified according to environmental controls. These results indicate stronger vertical <span class="hlt">velocity</span> profile behaviors under higher convective available potential energy and lower low-level moisture conditions. Sharp contrasts in convective area fraction and mass flux profiles are most pronounced when retrievals are segregated according to Amazonian wet and dry season conditions. During this deployment, wet season regimes favored higher domain mass flux profiles, attributed to more frequent convection that offsets weaker average convective cell vertical <span class="hlt">velocities</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29906713','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29906713"><span>Determination of elastic moduli from measured acoustic <span class="hlt">velocities</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brown, J Michael</p> <p>2018-06-01</p> <p>Methods are evaluated in solution of the inverse problem associated with determination of elastic moduli for crystals of arbitrary symmetry from elastic wave <span class="hlt">velocities</span> measured in many crystallographic directions. A package of MATLAB functions provides a robust and flexible environment for analysis of ultrasonic, Brillouin, or Impulsive Stimulated Light Scattering datasets. Three inverse algorithms are considered: the <span class="hlt">gradient</span>-based methods of Levenberg-Marquardt and Backus-Gilbert, and a non-<span class="hlt">gradient</span>-based (Nelder-Mead) simplex approach. Several data types are considered: body wave <span class="hlt">velocities</span> alone, surface wave <span class="hlt">velocities</span> plus a side constraint on X-ray-diffraction-based axes compressibilities, or joint body and surface wave <span class="hlt">velocities</span>. The numerical algorithms are validated through comparisons with prior published results and through analysis of synthetic datasets. Although all approaches succeed in finding low-misfit solutions, the Levenberg-Marquardt method consistently demonstrates effectiveness and computational efficiency. However, linearized <span class="hlt">gradient</span>-based methods, when applied to a strongly non-linear problem, may not adequately converge to the global minimum. The simplex method, while slower, is less susceptible to being trapped in local misfit minima. A "multi-start" strategy (initiate searches from more than one initial guess) provides better assurance that global minima have been located. Numerical estimates of parameter uncertainties based on Monte Carlo simulations are compared to formal uncertainties based on covariance calculations. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.S23C2737S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.S23C2737S"><span><span class="hlt">Velocity</span> Structure of the Iran Region Using Seismic and Gravity <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Syracuse, E. M.; Maceira, M.; Phillips, W. S.; Begnaud, M. L.; Nippress, S. E. J.; Bergman, E.; Zhang, H.</p> <p>2015-12-01</p> <p>We present a 3D Vp and Vs model of Iran generated using a joint inversion of body wave travel times, Rayleigh wave dispersion curves, and high-wavenumber filtered Bouguer gravity <span class="hlt">observations</span>. Our work has two main goals: 1) To better understand the tectonics of a prominent example of continental collision, and 2) To assess the improvements in earthquake location possible as a result of joint inversion. The body wave dataset is mainly derived from previous work on location calibration and includes the first-arrival P and S phases of 2500 earthquakes whose initial locations qualify as GT25 or better. The surface wave dataset consists of Rayleigh wave group <span class="hlt">velocity</span> measurements for regional earthquakes, which are inverted for a suite of period-dependent Rayleigh wave <span class="hlt">velocity</span> maps prior to inclusion in the joint inversion for body wave <span class="hlt">velocities</span>. We use gravity anomalies derived from the global gravity model EGM2008. To avoid mapping broad, possibly dynamic features in the gravity field intovariations in density and body wave <span class="hlt">velocity</span>, we apply a high-pass wavenumber filter to the gravity measurements. We use a simple, approximate relationship between density and <span class="hlt">velocity</span> so that the three datasets may be combined in a single inversion. The final optimized 3D Vp and Vs model allows us to explore how multi-parameter tomography addresses crustal heterogeneities in areas of limited coverage and improves travel time predictions. We compare earthquake locations from our models to independent locations obtained from InSAR analysis to assess the improvement in locations derived in a joint-inversion model in comparison to those derived in a more traditional body-wave-only <span class="hlt">velocity</span> model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900065465&hterms=1601&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231601','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900065465&hterms=1601&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231601"><span>Optical <span class="hlt">observations</span> on the CRIT-II Critical Ionization <span class="hlt">Velocity</span> Experiment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Stenbaek-Nielsen, H. C.; Wescott, E. M.; Haerendel, G.; Valenzuela, A.</p> <p>1990-01-01</p> <p>A rocket borne Critical Ionization <span class="hlt">Velocity</span> (CIV0 experiment was carried out from Wallops Island at dusk on May 4, 1989. Two barium shaped charges were released below the solar terminator (to prevent photoionization) at altitudes near 400 km. The ambient ionospheric electron density was 50,000/cu cm. The neutral barium jet was directed upward and at an angle of nominally 45 degrees to B which gives approximately 3 x 10 to the 23rd neutrals with super critical <span class="hlt">velocity</span>. Ions created by a CIV process in the region of the neutral jet would travel up along B into sunlight where they can be detected optically. Well defined ion clouds (max. brightness 750 R) were <span class="hlt">observed</span> in both releases. An ionization rate of 0.8 percent/sec (125 sec ionization time constant) can account for the <span class="hlt">observed</span> ion cloud near the release field line, but the ionization rate falls off with increasing distance from the release. It is concluded that a CIV process was present in the neutral jet out to about 50 km from the release, which is significantly further than allowed by current theories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25d2905H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25d2905H"><span>Particle simulation of electromagnetic emissions from electrostatic instability driven by an electron ring beam on the density <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Horký, Miroslav; Omura, Yoshiharu; Santolík, Ondřej</p> <p>2018-04-01</p> <p>This paper presents the wave mode conversion between electrostatic and electromagnetic waves on the plasma density <span class="hlt">gradient</span>. We use 2-D electromagnetic code KEMPO2 implemented with the generation of density <span class="hlt">gradient</span> to simulate such a conversion process. In the dense region, we use ring beam instability to generate electron Bernstein waves and we study the temporal evolution of wave spectra, <span class="hlt">velocity</span> distributions, Poynting flux, and electric and magnetic energies to <span class="hlt">observe</span> the wave mode conversion. Such a conversion process can be a source of electromagnetic emissions which are routinely measured by spacecraft on the plasmapause density <span class="hlt">gradient</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.P51B2061D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.P51B2061D"><span>From pebbles to dust: experiments to <span class="hlt">observe</span> low-<span class="hlt">velocity</span> collisional outcomes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dove, A.; Jorges, J.; Colwell, J. E.</p> <p>2015-12-01</p> <p>Particle size evolution in planetary ring systems can be driven by collisions at relatively low <span class="hlt">velocities</span> (<1 m/s) occurring between objects with a range of sizes from very fine dust to decimeter-sized objects. In these complex systems, collisions between centimeter-sized objects may result in particle growth by accretion, rebounding, or erosive processes that result in the production of additional smaller particles. The outcomes of these collisions are dependent on factors such as collisional energy, particle size, and particle morphology. Numerical simulations are limited by a need to understand these collisional parameters over a range of conditions. We present the results of a sequence of laboratory experiments designed to explore collisions over a range of these parameters. We are able to <span class="hlt">observe</span> low-<span class="hlt">velocity</span> collisions by conducting experiments in vacuum chambers in our 0.8-sec drop tower apparatus. Initial experiments utilize a variety of impacting spheres, including glass, Teflon, aluminum, stainless steel, and brass. These spheres are either used in their natural state or are "mantled" - coated with a few-mm thick layer of a cohesive powder. A high-speed, high-resolution video camera is used to record the motion of the colliding bodies. These videos are then processed and we track the particles to determine impactor speeds before and after collision and the collisional outcome; in the case of the mantled impactors, we can assess how much of the powder was released in the collision. We also determine how the coefficient of restitution varies as a function of material type, morphology, and impact <span class="hlt">velocity</span>. Impact <span class="hlt">velocities</span> range from about 20-60 cm/s, and we <span class="hlt">observe</span> that mantling of particles significantly reduces their coefficients of restitution. These results will contribute to an empirical model of collisional outcomes that can help refine our understanding of dusty ring system collisional evolution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMOS11A1256H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMOS11A1256H"><span>Surfzone alongshore advective accelerations: <span class="hlt">observations</span> and modeling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hansen, J.; Raubenheimer, B.; Elgar, S.</p> <p>2014-12-01</p> <p>The sources, magnitudes, and impacts of non-linear advective accelerations on alongshore surfzone currents are investigated with <span class="hlt">observations</span> and a numerical model. Previous numerical modeling results have indicated that advective accelerations are an important contribution to the alongshore force balance, and are required to understand spatial variations in alongshore currents (which may result in spatially variable morphological change). However, most prior <span class="hlt">observational</span> studies have neglected advective accelerations in the alongshore force balance. Using a numerical model (Delft3D) to predict optimal sensor locations, a dense array of 26 colocated current meters and pressure sensors was deployed between the shoreline and 3-m water depth over a 200 by 115 m region near Duck, NC in fall 2013. The array included 7 cross- and 3 alongshore transects. Here, <span class="hlt">observational</span> and numerical estimates of the dominant forcing terms in the alongshore balance (pressure and radiation-stress <span class="hlt">gradients</span>) and the advective acceleration terms will be compared with each other. In addition, the numerical model will be used to examine the force balance, including sources of <span class="hlt">velocity</span> <span class="hlt">gradients</span>, at a higher spatial resolution than possible with the instrument array. Preliminary numerical results indicate that at O(10-100 m) alongshore scales, bathymetric variations and the ensuing alongshore variations in the wave field and subsequent forcing are the dominant sources of the modeled <span class="hlt">velocity</span> <span class="hlt">gradients</span> and advective accelerations. Additional simulations and analysis of the <span class="hlt">observations</span> will be presented. Funded by NSF and ASDR&E.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.792a2092B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.792a2092B"><span>Multiple <span class="hlt">velocity</span> encoding in the phase of an MRI signal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benitez-Read, E. E.</p> <p>2017-01-01</p> <p>The measurement of fluid <span class="hlt">velocity</span> by encoding it in the phase of a magnetic resonance imaging (MRI) signal could allow the discrimination of the stationary spins signals from those of moving spins. This results in a wide variety of applications i.e. in medicine, in order to obtain more than angiograms, blood <span class="hlt">velocity</span> images of veins, arteries and other vessels without having static tissue perturbing the signal of fluid in motion. The work presented in this paper is a theoretical analysis of some novel methods for multiple fluid <span class="hlt">velocity</span> encoding in the phase of an MRI signal. These methods are based on a tripolar <span class="hlt">gradient</span> (TPG) and can be an alternative to the conventional methods based on a bipolar <span class="hlt">gradient</span> (BPG) and could be more suitable for multiple <span class="hlt">velocity</span> encoding in the phase of an MRI signal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..1513300H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..1513300H"><span>Modeling and comparative study of fluid <span class="hlt">velocities</span> in heterogeneous rocks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hingerl, Ferdinand F.; Romanenko, Konstantin; Pini, Ronny; Balcom, Bruce; Benson, Sally</p> <p>2013-04-01</p> <p>Detailed knowledge of the distribution of effective porosity and fluid <span class="hlt">velocities</span> in heterogeneous rock samples is crucial for understanding and predicting spatially resolved fluid residence times and kinetic reaction rates of fluid-rock interactions. The applicability of conventional MRI techniques to sedimentary rocks is limited by internal magnetic field <span class="hlt">gradients</span> and short spin relaxation times. The approach developed at the UNB MRI Centre combines the 13-interval Alternating-Pulsed-<span class="hlt">Gradient</span> Stimulated-Echo (APGSTE) scheme and three-dimensional Single Point Ramped Imaging with T1 Enhancement (SPRITE). These methods were designed to reduce the errors due to effects of background <span class="hlt">gradients</span> and fast transverse relaxation. SPRITE is largely immune to time-evolution effects resulting from background <span class="hlt">gradients</span>, paramagnetic impurities and chemical shift. Using these techniques quantitative 3D porosity maps as well as single-phase fluid <span class="hlt">velocity</span> fields in sandstone core samples were measured. Using a new Magnetic Resonance Imaging technique developed at the MRI Centre at UNB, we created 3D maps of porosity distributions as well as single-phase fluid <span class="hlt">velocity</span> distributions of sandstone rock samples. Then, we evaluated the applicability of the Kozeny-Carman relationship for modeling measured fluid <span class="hlt">velocity</span> distributions in sandstones samples showing meso-scale heterogeneities using two different modeling approaches. The MRI maps were used as reference points for the modeling approaches. For the first modeling approach, we applied the Kozeny-Carman relationship to the porosity distributions and computed respective permeability maps, which in turn provided input for a CFD simulation - using the Stanford CFD code GPRS - to compute averaged <span class="hlt">velocity</span> maps. The latter were then compared to the measured <span class="hlt">velocity</span> maps. For the second approach, the measured <span class="hlt">velocity</span> distributions were used as input for inversely computing permeabilities using the GPRS CFD code. The computed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24926506','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24926506"><span>A study on Rayleigh wave dispersion in bone according to Mindlin's Form II <span class="hlt">gradient</span> elasticity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vavva, Maria G; Gergidis, Leonidas N; Protopappas, Vasilios C; Charalambopoulos, Antonios; Polyzos, Demosthenes; Fotiadis, Dimitrios I</p> <p>2014-05-01</p> <p>The classical elasticity cannot effectively describe bone's mechanical behavior since only homogeneous media and local stresses are assumed. Additionally, it cannot predict the dispersive nature of the Rayleigh wave which has been reported in experimental studies and was also demonstrated in a previous computational study by adopting Mindlin's Form II <span class="hlt">gradient</span> elasticity. In this work Mindlin's theory is employed to analytically determine the dispersion of Rayleigh waves in a strain <span class="hlt">gradient</span> elastic half-space. An isotropic semi-infinite space is considered with properties equal to those of bone and dynamic behavior suffering from microstructural effects. Microstructural effects are considered by incorporating four intrinsic parameters in the stress analysis. The results are presented in the form of group and phase <span class="hlt">velocity</span> dispersion curves and compared with existing computational results and semi-analytical curves calculated for a simpler case of Rayleigh waves in dipolar <span class="hlt">gradient</span> elastic half-spaces. Comparisons are also performed with the <span class="hlt">velocity</span> of the first-order antisymmetric mode propagating in a dipolar plate so as to <span class="hlt">observe</span> the Rayleigh asymptotic behavior. It is shown that Mindlin's Form II <span class="hlt">gradient</span> elasticity can effectively describe the dispersive nature of Rayleigh waves. This study could be regarded as a step toward the ultrasonic characterization of bone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017HESS...21.3221N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017HESS...21.3221N"><span>Slope-<span class="hlt">velocity</span> equilibrium and evolution of surface roughness on a stony hillslope</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nearing, Mark A.; Polyakov, Viktor O.; Nichols, Mary H.; Hernandez, Mariano; Li, Li; Zhao, Ying; Armendariz, Gerardo</p> <p>2017-06-01</p> <p>Slope-<span class="hlt">velocity</span> equilibrium is hypothesized as a state that evolves naturally over time due to the interaction between overland flow and surface morphology, wherein steeper areas develop a relative increase in physical and hydraulic roughness such that flow <span class="hlt">velocity</span> is a unique function of overland flow rate independent of slope <span class="hlt">gradient</span>. This study tests this hypothesis under controlled conditions. Artificial rainfall was applied to 2 m by 6 m plots at 5, 12, and 20 % slope <span class="hlt">gradients</span>. A series of simulations were made with two replications for each treatment with measurements of runoff rate, <span class="hlt">velocity</span>, rock cover, and surface roughness. <span class="hlt">Velocities</span> measured at the end of each experiment were a unique function of discharge rates, independent of slope <span class="hlt">gradient</span> or rainfall intensity. Physical surface roughness was greater at steeper slopes. The data clearly showed that there was no unique hydraulic coefficient for a given slope, surface condition, or rainfall rate, with hydraulic roughness greater at steeper slopes and lower intensities. This study supports the hypothesis of slope-<span class="hlt">velocity</span> equilibrium, implying that use of hydraulic equations, such as Chezy and Manning, in hillslope-scale runoff models is problematic because the coefficients vary with both slope and rainfall intensity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70037715','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70037715"><span>Evaluation of Maryland abutment scour equation through selected threshold <span class="hlt">velocity</span> methods</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Benedict, S.T.</p> <p>2010-01-01</p> <p>The U.S. Geological Survey, in cooperation with the Maryland State Highway Administration, used field measurements of scour to evaluate the sensitivity of the Maryland abutment scour equation to the critical (or threshold) <span class="hlt">velocity</span> variable. Four selected methods for estimating threshold <span class="hlt">velocity</span> were applied to the Maryland abutment scour equation, and the predicted scour to the field measurements were compared. Results indicated that performance of the Maryland abutment scour equation was sensitive to the threshold <span class="hlt">velocity</span> with some threshold <span class="hlt">velocity</span> methods producing better estimates of predicted scour than did others. In addition, results indicated that regional stream characteristics can affect the performance of the Maryland abutment scour equation with moderate-<span class="hlt">gradient</span> streams performing differently from low-<span class="hlt">gradient</span> streams. On the basis of the findings of the investigation, guidance for selecting threshold <span class="hlt">velocity</span> methods for application to the Maryland abutment scour equation are provided, and limitations are noted.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Sci...354..312T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Sci...354..312T"><span>Dynamic creation and evolution of <span class="hlt">gradient</span> nanostructure in single-crystal metallic microcubes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thevamaran, Ramathasan; Lawal, Olawale; Yazdi, Sadegh; Jeon, Seog-Jin; Lee, Jae-Hwang; Thomas, Edwin L.</p> <p>2016-10-01</p> <p>We demonstrate the dynamic creation and subsequent static evolution of extreme <span class="hlt">gradient</span> nanograined structures in initially near-defect-free single-crystal silver microcubes. Extreme nanostructural transformations are imposed by high strain rates, strain <span class="hlt">gradients</span>, and recrystallization in high-<span class="hlt">velocity</span> impacts of the microcubes against an impenetrable substrate. We synthesized the silver microcubes in a bottom-up seed-growth process and use an advanced laser-induced projectile impact testing apparatus to selectively launch them at supersonic <span class="hlt">velocities</span> (~400 meters per second). Our study provides new insights into the fundamental deformation mechanisms and the effects of crystal and sample-shape symmetries resulting from high-<span class="hlt">velocity</span> impacts. The nanostructural transformations produced in our experiments show promising pathways to developing <span class="hlt">gradient</span> nanograined metals for engineering applications requiring both high strength and high toughness—for example, in structural components of aircraft and spacecraft.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10139E..0JR','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10139E..0JR"><span>Lateral <span class="hlt">velocity</span> estimation bias due to beamforming delay errors (Conference Presentation)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodriguez-Molares, Alfonso; Fadnes, Solveig; Swillens, Abigail; Løvstakken, Lasse</p> <p>2017-03-01</p> <p>An artefact has recently been reported [1,2] in the estimation of the lateral blood <span class="hlt">velocity</span> using speckle tracking. This artefact shows as a net <span class="hlt">velocity</span> bias in presence of strong spatial <span class="hlt">velocity</span> <span class="hlt">gradients</span> such as those that occur at the edges of the filling jets in the heart. Even though this artifact has been found both in vitro and in simulated data, its causes are still undescribed. Here we demonstrate that a potential source of this artefact can be traced to smaller errors in the beamforming setup. By inserting a small offset in the beamforming delay, one can artificially create a net lateral movement in the speckle in areas of high <span class="hlt">velocity</span> <span class="hlt">gradient</span>. That offset does not have a strong impact in the image quality and can easily go undetected.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.467.2787H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.467.2787H"><span>Not a Copernican <span class="hlt">observer</span>: biased peculiar <span class="hlt">velocity</span> statistics in the local Universe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hellwing, Wojciech A.; Nusser, Adi; Feix, Martin; Bilicki, Maciej</p> <p>2017-05-01</p> <p>We assess the effect of the local large-scale structure on the estimation of two-point statistics of the <span class="hlt">observed</span> radial peculiar <span class="hlt">velocities</span> of galaxies. A large N-body simulation is used to examine these statistics from the perspective of random <span class="hlt">observers</span> as well as 'Local Group-like' <span class="hlt">observers</span> conditioned to reside in an environment resembling the <span class="hlt">observed</span> Universe within 20 Mpc. The local environment systematically distorts the shape and amplitude of <span class="hlt">velocity</span> statistics with respect to ensemble-averaged measurements made by a Copernican (random) <span class="hlt">observer</span>. The Virgo cluster has the most significant impact, introducing large systematic deviations in all the statistics. For a simple 'top-hat' selection function, an idealized survey extending to ˜160 h-1 Mpc or deeper is needed to completely mitigate the effects of the local environment. Using shallower catalogues leads to systematic deviations of the order of 50-200 per cent depending on the scale considered. For a flat redshift distribution similar to the one of the CosmicFlows-3 survey, the deviations are even more prominent in both the shape and amplitude at all separations considered (≲100 h-1 Mpc). Conclusions based on statistics calculated without taking into account the impact of the local environment should be revisited.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995JGR...100.6143K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995JGR...100.6143K"><span>A baseline for upper crustal <span class="hlt">velocity</span> variations along the East Pacific Rise at 13 deg N</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kappus, Mary E.; Harding, Alistair J.; Orcutt, John A.</p> <p>1995-04-01</p> <p>A wide aperture profile of the East Pacific Rise at 13 deg N provides data necessary to make a high-resolution seismic <span class="hlt">velocity</span> profile of the uppermost crust along a 52-km segment of ridge crest. Automated and objective processing steps, including r-p analysis and waveform inversion, allow the construction of models in a consistent way so that comparisons are meaningful. A continuous profile is synthesized from 70 independent one-dimensional models spaced at 750-km intervals along the ridge. The resulting seismic <span class="hlt">velocity</span> structure of the top 500 m of crust is remarkable in its lack of variability. The main features are a thin low-<span class="hlt">velocity</span> layer 2A at the top with a steep <span class="hlt">gradient</span> to layer 2B. The seafloor <span class="hlt">velocity</span> is nearly constant at 2.45 km/s +/- 3% along the entire ridge. The <span class="hlt">velocity</span> at the top of layer 2B is 5.0 km/s +/- 10%. The depth to the 4 km/s isovelocity contour within layer 2A is 130 +/- 20 m from 13 deg to 13 deg 20 min N, north of which it increases to 180 m. The increase in thickness is coincident with a deviation from axial linearity (DEVAL) noted by both a slight change in axis depth and orientation and in geochemistry. The waveform inversion, providing more details plus <span class="hlt">velocity</span> <span class="hlt">gradient</span> information, shows a layer 2A with about 80 m of constant-<span class="hlt">velocity</span> material underlain by 150 m of high <span class="hlt">velocity</span> <span class="hlt">gradient</span> material, putting the base of layer 2A at approximately 230 m depth south of 13 deg 20 min N and about 50 m thicker north of the DEVAL. The overall lack of variability, combined with other recent measurements of layer 2A thickness along and near the axis, indicates that the thickness of volcanic extrusives is controlled not by levels of volcanic productivity, but the dynamics of emplacement. The homogeneity along axis also provides a baseline of inherent variability in crustal structure of about 10% against which other <span class="hlt">observed</span> variations in similar regimes can be compared.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014Ap%26SS.351..289K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014Ap%26SS.351..289K"><span>A comparison between <span class="hlt">observed</span> and analytical <span class="hlt">velocity</span> dispersion profiles of 20 nearby galaxy clusters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khan, Mohammad S.; Abdullah, Mohamed H.; Ali, Gamal B.</p> <p>2014-05-01</p> <p>We derive analytical expression for the <span class="hlt">velocity</span> dispersion of galaxy clusters, using the statistical mechanical approach. We compare the <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersion profiles for 20 nearby ( z≤0.1) galaxy clusters with the analytical ones. It is interesting to find that the analytical results closely match with the <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersion profiles only if the presence of the diffuse matter in clusters is taken into consideration. This takes us to introduce a new approach to detect the ratio of diffuse mass, M diff , within a galaxy cluster. For the present sample, the ratio f= M diff / M, where M the cluster's total mass is found to has an average value of 45±12 %. This leads us to the result that nearly 45 % of the cluster mass is impeded outside the galaxies, while around 55 % of the cluster mass is settled in the galaxies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeoJI.204.1490G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeoJI.204.1490G"><span>Field <span class="hlt">observations</span> of seismic <span class="hlt">velocity</span> changes caused by shaking-induced damage and healing due to mesoscopic nonlinearity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gassenmeier, M.; Sens-Schönfelder, C.; Eulenfeld, T.; Bartsch, M.; Victor, P.; Tilmann, F.; Korn, M.</p> <p>2016-03-01</p> <p>To investigate temporal seismic <span class="hlt">velocity</span> changes due to earthquake related processes and environmental forcing in Northern Chile, we analyse 8 yr of ambient seismic noise recorded by the Integrated Plate Boundary Observatory Chile (IPOC). By autocorrelating the ambient seismic noise field measured on the vertical components, approximations of the Green's functions are retrieved and <span class="hlt">velocity</span> changes are measured with Coda Wave Interferometry. At station PATCX, we <span class="hlt">observe</span> seasonal changes in seismic <span class="hlt">velocity</span> caused by thermal stress as well as transient <span class="hlt">velocity</span> reductions in the frequency range of 4-6 Hz. Sudden <span class="hlt">velocity</span> drops occur at the time of mostly earthquake-induced ground shaking and recover over a variable period of time. We present an empirical model that describes the seismic <span class="hlt">velocity</span> variations based on continuous <span class="hlt">observations</span> of the local ground acceleration. The model assumes that not only the shaking of large earthquakes causes <span class="hlt">velocity</span> drops, but any small vibrations continuously induce minor <span class="hlt">velocity</span> variations that are immediately compensated by healing in the steady state. We show that the shaking effect is accumulated over time and best described by the integrated envelope of the ground acceleration over the discretization interval of the <span class="hlt">velocity</span> measurements, which is one day. In our model, the amplitude of the <span class="hlt">velocity</span> reduction as well as the recovery time are proportional to the size of the excitation. This model with two free scaling parameters fits the data of the shaking induced <span class="hlt">velocity</span> variation in remarkable detail. Additionally, a linear trend is <span class="hlt">observed</span> that might be related to a recovery process from one or more earthquakes before our measurement period. A clear relationship between ground shaking and induced <span class="hlt">velocity</span> reductions is not visible at other stations. We attribute the outstanding sensitivity of PATCX to ground shaking and thermal stress to the special geological setting of the station, where the subsurface material</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvF...3c4606V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvF...3c4606V"><span>Determination of wall shear stress from mean <span class="hlt">velocity</span> and Reynolds shear stress profiles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Volino, Ralph J.; Schultz, Michael P.</p> <p>2018-03-01</p> <p>An analytical method is presented for determining the Reynolds shear stress profile in steady, two-dimensional wall-bounded flows using the mean streamwise <span class="hlt">velocity</span>. The method is then utilized with experimental data to determine the local wall shear stress. The procedure is applicable to flows on smooth and rough surfaces with arbitrary pressure <span class="hlt">gradients</span>. It is based on the streamwise component of the boundary layer momentum equation, which is transformed into inner coordinates. The method requires <span class="hlt">velocity</span> profiles from at least two streamwise locations, but the formulation of the momentum equation reduces the dependence on streamwise <span class="hlt">gradients</span>. The method is verified through application to laminar flow solutions and turbulent DNS results from both zero and nonzero pressure <span class="hlt">gradient</span> boundary layers. With strong favorable pressure <span class="hlt">gradients</span>, the method is shown to be accurate for finding the wall shear stress in cases where the Clauser fit technique loses accuracy. The method is then applied to experimental data from the literature from zero pressure <span class="hlt">gradient</span> studies on smooth and rough walls, and favorable and adverse pressure <span class="hlt">gradient</span> cases on smooth walls. Data from very near the wall are not required for determination of the wall shear stress. Wall friction <span class="hlt">velocities</span> obtained using the present method agree with those determined in the original studies, typically to within 2%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970021344','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970021344"><span>Structures in Ionospheric Number Density and <span class="hlt">Velocity</span> Associated with Polar Cap Ionization Patches</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kivanc, O.; Heelis, R. A.</p> <p>1997-01-01</p> <p>Spectral characteristics of polar cap F region irregularities on large density <span class="hlt">gradients</span> associated with polar ionization patches are studied using in situ measurements made by the Dynamics Explorer 2 (DE 2) spacecraft. The 18 patches studied in this paper were identified by the algorithm introduced by Coley and Heelis, and they were encountered during midnight-noon passes of the spacecraft. Density and <span class="hlt">velocity</span> spectra associated with these antisunward convecting patches are analyzed in detail. <span class="hlt">Observations</span> indicate the presence of structure on most patches regardless of the distance between the patch and the cusp where they are believed to develop. Existence of structure on both leading and trailing edges is established when such edges exist. Results, which show no large dependence of Delta N/N power on the sign of the edge <span class="hlt">gradient</span> del N, do not allow the identification of leading and trailing edges of the patch. The Delta N/N is an increasing function of <span class="hlt">gradient</span> del N regardless of the sign of the <span class="hlt">gradient</span>. The correlation between Delta N/N and Delta V is generally poor, but for a given intensity in Delta V, Delta N/N maximizes in regions of large <span class="hlt">gradients</span> in N. There is evidence for the presence of unstructured patches that seem to co-exist with unstructured horizontal <span class="hlt">velocities</span>. Slightly smaller spectral indices for trailing edges support the presence of the E X B drift instability. Although this instability is found to be operating in some cases, results suggest that stirring may be a significant contributor to kilometer-size structures in the polar cap.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..DPPBP8007K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..DPPBP8007K"><span>Effects of density <span class="hlt">gradient</span> caused by multi-pulsing CHI on two-fluid flowing equilibria of spherical torus plasmas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kanki, T.; Nagata, M.</p> <p>2014-10-01</p> <p>Two-fluid dynamo relaxation is examined to understand sustainment mechanism of spherical torus (ST) plasmas by multi-pulsing CHI (M-CHI) in the HIST device. The steeper density <span class="hlt">gradient</span> between the central open flux column (OFC) and closed flux regions by applying the second CHI pulse is <span class="hlt">observed</span> to cause not only the <emph type="bold-italic">E</emph> × <emph type="bold-italic">B</emph> drift but also the ion diamagnetic drift, leading the two-fluid dynamo. The purpose of this study is to investigate the effects of the steep change in the density <span class="hlt">gradient</span> on the ST equilibria by using the two-fluid equilibrium calculations. The toroidal magnetic field becomes from a diamagnetic to a paramagnetic profile in the closed flux region while it remains a diamagnetic profile in the OFC region. The toroidal ion flow <span class="hlt">velocity</span> is increased from negative to positive values in the closed flux region. Here, the negative ion flow <span class="hlt">velocity</span> is the opposite direction to the toroidal current. The poloidal ion flow <span class="hlt">velocity</span> between the OFC and closed flux regions is increased, because the ion diamagnetic drift <span class="hlt">velocity</span> is changed in the same direction as the <emph type="bold-italic">E</emph> × <emph type="bold-italic">B</emph> drift <span class="hlt">velocity</span> through the steeper ion pressure <span class="hlt">gradient</span>. As a result, the strong shear flow and the paramagnetic toroidal field are generated in the closed flux region. Here, the ion flow <span class="hlt">velocity</span> is the same direction as the poloidal current. The radial electric field shear between the OFC and closed flux regions is enhanced due to the strong dependence on the magnetic force through the interaction of toroidal ion flow <span class="hlt">velocity</span> and axial magnetic field. The two-fluid effect is significant there due to the ion diamagnetic effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21147123-response-dynamics-bluff-body-stabilized-conical-premixed-turbulent-flames-spatial-mixture-gradients','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21147123-response-dynamics-bluff-body-stabilized-conical-premixed-turbulent-flames-spatial-mixture-gradients"><span>Response dynamics of bluff-body stabilized conical premixed turbulent flames with spatial mixture <span class="hlt">gradients</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Chaudhuri, Swetaprovo; Cetegen, Baki M.</p> <p>2009-03-15</p> <p>Response of bluff-body stabilized conical turbulent premixed flames was experimentally studied for a range of excitation frequencies (10-400 Hz), mean flow <span class="hlt">velocities</span> (5, 10 and 15 m/s) and three different spatial mixture distributions (uniform, inner and outer enrichment). Upstream excitation was provided by a loudspeaker producing <span class="hlt">velocity</span> oscillation amplitudes of about 8% of the mean flow <span class="hlt">velocity</span>. Flame response was detected by a photomultiplier <span class="hlt">observing</span> the CH{sup *} emission from the flame. The studied turbulent flames exhibited transfer function characteristics of a low-pass filter with a cutoff Strouhal number between 0.08 and 0.12. The amplification factors at low frequencies rangedmore » from 2 to 20 and generally increased for mean flow <span class="hlt">velocities</span> from 5 to 15 m/s. The highest levels of amplification were found for the outer mixture enrichment followed in decreasing order by uniform and inner mixture <span class="hlt">gradient</span> cases. The high levels of flame response for the outer enrichment case were attributed to the enhanced flame-vortex interaction in outer jet shear layer. At high excitation levels (u{sup '}/U{sub m}{approx}0.3) for U{sub m}=5 m/ s where non-linear flame response is expected, the flame exhibited a reduced amplitude response in the frequency range between 40 and 100 Hz for the uniform and outer equivalence ratio <span class="hlt">gradient</span> cases and no discernible effect for the inner equivalence ratio <span class="hlt">gradient</span>. In all cases, transfer function phase was found to vary linearly with excitation frequency. Finally, a relationship between the amplitude characteristics of the bluff-body wake transfer function and flame blowoff equivalence ratio was presented. (author)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPD....4830301Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPD....4830301Y"><span>Using <span class="hlt">observations</span> of slipping <span class="hlt">velocities</span> to test the hypothesis that reconnection heats the active region corona</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Kai; Longcope, Dana; Guo, Yang; Ding, Mingde</p> <p>2017-08-01</p> <p>Numerous proposed coronal heating mechanisms have invoked magnetic reconnection in some role. Testing such a mechanism requires a method of measuring magnetic reconnection coupled with a prediction of the heat delivered by reconnection at the <span class="hlt">observed</span> rate. In the absence of coronal reconnection, field line footpoints move at the same <span class="hlt">velocity</span> as the plasma they find themselves in. The rate of coronal reconnection is therefore related to any discrepancy <span class="hlt">observed</span> between footpoint motion and that of the local plasma — so-called slipping motion. We propose a novel method to measure this <span class="hlt">velocity</span> discrepancy by combining a sequence of non-linear force-free field extrapolations with maps of photospheric <span class="hlt">velocity</span>. We obtain both from a sequence of vector magnetograms of an active region (AR). We then propose a method of computing the coronal heating produced under the assumption the <span class="hlt">observed</span> slipping <span class="hlt">velocity</span> was due entirely to coronal reconnection. This heating rate is used to predict density and temperature at points along an equilibrium loop. This, in turn, is used to synthesize emission in EUV and SXR bands. We perform this analysis using a sequence of HMI vector magnetograms of a particular AR and compare synthesized images to <span class="hlt">observations</span> of the same AR made by SDO. We also compare differential emission measure inferred from those <span class="hlt">observations</span> to that of the modeled corona.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27846562','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27846562"><span>Dynamic creation and evolution of <span class="hlt">gradient</span> nanostructure in single-crystal metallic microcubes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thevamaran, Ramathasan; Lawal, Olawale; Yazdi, Sadegh; Jeon, Seog-Jin; Lee, Jae-Hwang; Thomas, Edwin L</p> <p>2016-10-21</p> <p>We demonstrate the dynamic creation and subsequent static evolution of extreme <span class="hlt">gradient</span> nanograined structures in initially near-defect-free single-crystal silver microcubes. Extreme nanostructural transformations are imposed by high strain rates, strain <span class="hlt">gradients</span>, and recrystallization in high-<span class="hlt">velocity</span> impacts of the microcubes against an impenetrable substrate. We synthesized the silver microcubes in a bottom-up seed-growth process and use an advanced laser-induced projectile impact testing apparatus to selectively launch them at supersonic <span class="hlt">velocities</span> (~400 meters per second). Our study provides new insights into the fundamental deformation mechanisms and the effects of crystal and sample-shape symmetries resulting from high-<span class="hlt">velocity</span> impacts. The nanostructural transformations produced in our experiments show promising pathways to developing <span class="hlt">gradient</span> nanograined metals for engineering applications requiring both high strength and high toughness-for example, in structural components of aircraft and spacecraft. Copyright © 2016, American Association for the Advancement of Science.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAESc.134..231M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAESc.134..231M"><span>Group <span class="hlt">velocity</span> dispersion characteristics and one-dimensional regional shear <span class="hlt">velocity</span> structure of the eastern Indian craton</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mandal, Prantik</p> <p>2017-02-01</p> <p>In the past three years, a semi-permanent network of fifteen 3-component broadband seismographs has become operational in the eastern Indian shield region occupying the Archean (∼2.5-3.6 Ga) Singhbhum-Odisha craton (SOC) and the Proterozoic (∼1.0-2.5 Ga) Chotanagpur Granitic Gneissic terrane (CGGT). The reliable and accurate broadband data for the recent 2015 Nepal earthquake sequence from 10 broadband stations of this network enabled us to estimate the group <span class="hlt">velocity</span> dispersion characteristics and one-dimensional regional shear <span class="hlt">velocity</span> structure of the region. First, we measure fundamental mode Rayleigh- and Love-wave group <span class="hlt">velocity</span> dispersion curves in the period range of 7-70 s and then invert these curves to estimate the crustal and upper mantle structure below the eastern Indian craton (EIC). We <span class="hlt">observe</span> that group <span class="hlt">velocities</span> of Rayleigh and Love waves in SOC are relatively high in comparison to those of CGGT. This could be attributed to a relatively mafic-rich crust-mantle structure in SOC resulting from two episodes of magmatism associated with the 1.6 Ga Dalma and ∼117 Ma Rajmahal volcanisms. The best model for the EIC from the present study is found to be a two-layered crust, with a 14-km thick upper-crust (UC) of average shear <span class="hlt">velocity</span> (Vs) of 3.0 km/s and a 26-km thick lower-crust (LC) of average Vs of 3.6 km/s. The present study detects a sharp drop in Vs (∼-2 to 3%) at 120-260 km depths, underlying the EIC, representing the probable seismic lithosphere-asthenosphere boundary (LAB) at 120 km depth. Such sharp fall in Vs below the LAB indicates a partially molten layer. Further, a geothermal <span class="hlt">gradient</span> extrapolated from the surface heat flow shows that such a <span class="hlt">gradient</span> would intercept the wet basalt solidus at 88-103 km depths, suggesting a 88-103 km thick thermal lithosphere below the EIC. This could also signal the presence of small amounts of partial melts. Thus, this 2-3% drop in Vs could be attributed to the presence of partial melts in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1344928','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1344928"><span>In Situ <span class="hlt">observation</span> of dark current emission in a high <span class="hlt">gradient</span> rf photocathode gun</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Shao, Jiahang; Shi, Jiaru; Antipov, Sergey P.</p> <p></p> <p>Undesirable electron field emission (also known as dark current) in high <span class="hlt">gradient</span> rf photocathode guns deteriorates the quality of the photoemission current and limits the operational <span class="hlt">gradient</span>. To improve the understanding of dark current emission, a high-resolution (~100 μm) dark current imaging experiment has been performed in an L-band photocathode gun operating at ~100 MV/m of surface <span class="hlt">gradient</span>. Scattered strong emission areas with high current have been <span class="hlt">observed</span> on the cathode. The field enhancement factor β of selected regions on the cathode has been measured. Finally, the postexaminations with scanning electron microscopy and white light interferometry reveal the origins ofmore » ~75% strong emission areas overlap with the spots where rf breakdown has occurred.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1344928-nbsp-situ-observation-dark-current-emission-high-gradient-rf-photocathode-gun','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1344928-nbsp-situ-observation-dark-current-emission-high-gradient-rf-photocathode-gun"><span>In Situ <span class="hlt">observation</span> of dark current emission in a high <span class="hlt">gradient</span> rf photocathode gun</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Shao, Jiahang; Shi, Jiaru; Antipov, Sergey P.; ...</p> <p>2016-08-15</p> <p>Undesirable electron field emission (also known as dark current) in high <span class="hlt">gradient</span> rf photocathode guns deteriorates the quality of the photoemission current and limits the operational <span class="hlt">gradient</span>. To improve the understanding of dark current emission, a high-resolution (~100 μm) dark current imaging experiment has been performed in an L-band photocathode gun operating at ~100 MV/m of surface <span class="hlt">gradient</span>. Scattered strong emission areas with high current have been <span class="hlt">observed</span> on the cathode. The field enhancement factor β of selected regions on the cathode has been measured. Finally, the postexaminations with scanning electron microscopy and white light interferometry reveal the origins ofmore » ~75% strong emission areas overlap with the spots where rf breakdown has occurred.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990PApGe.132..417P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990PApGe.132..417P"><span>Inverse kinematic problem for a random <span class="hlt">gradient</span> medium in geometric optics approximation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Petersen, N. V.</p> <p>1990-03-01</p> <p>Scattering at random inhomogeneities in a <span class="hlt">gradient</span> medium results in systematic deviations of the rays and travel times of refracted body waves from those corresponding to the deterministic <span class="hlt">velocity</span> component. The character of the difference depends on the parameters of the deterministic and random <span class="hlt">velocity</span> component. However, at great distances to the source, independently of the <span class="hlt">velocity</span> parameters (weakly or strongly inhomogeneous medium), the most probable depth of the ray turning point is smaller than that corresponding to the deterministic <span class="hlt">velocity</span> component, the most probable travel times also being lower. The relative uncertainty in the deterministic <span class="hlt">velocity</span> component, derived from the mean travel times using methods developed for laterally homogeneous media (for instance, the Herglotz-Wiechert method), is systematic in character, but does not exceed the contrast of <span class="hlt">velocity</span> inhomogeneities by magnitude. The <span class="hlt">gradient</span> of the deterministic <span class="hlt">velocity</span> component has a significant effect on the travel-time fluctuations. The variance at great distances to the source is mainly controlled by shallow inhomogeneities. The travel-time flucutations are studied only for weakly inhomogeneous media.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..577L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..577L"><span>High-resolution vertical <span class="hlt">velocities</span> and their power spectrum <span class="hlt">observed</span> with the MAARSY radar - Part 1: frequency spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Qiang; Rapp, Markus; Stober, Gunter; Latteck, Ralph</p> <p>2018-04-01</p> <p>The Middle Atmosphere Alomar Radar System (MAARSY) installed at the island of Andøya has been run for continuous probing of atmospheric winds in the upper troposphere and lower stratosphere (UTLS) region. In the current study, we present high-resolution wind measurements during the period between 2010 and 2013 with MAARSY. The spectral analysis applying the Lomb-Scargle periodogram method has been carried out to determine the frequency spectra of vertical wind <span class="hlt">velocity</span>. From a total of 522 days of <span class="hlt">observations</span>, the statistics of the spectral slope have been derived and show a dependence on the background wind conditions. It is a general feature that the <span class="hlt">observed</span> spectra of vertical <span class="hlt">velocity</span> during active periods (with wind <span class="hlt">velocity</span> > 10 m s-1) are much steeper than during quiet periods (with wind <span class="hlt">velocity</span> < 10 m s-1). The distribution of spectral slopes is roughly symmetric with a maximum at -5/3 during active periods, whereas a very asymmetric distribution with a maximum at around -1 is <span class="hlt">observed</span> during quiet periods. The slope profiles along altitudes reveal a significant height dependence for both conditions, i.e., the spectra become shallower with increasing altitudes in the upper troposphere and maintain roughly a constant slope in the lower stratosphere. With both wind conditions considered together the general spectra are obtained and their slopes are compared with the background horizontal winds. The comparisons show that the <span class="hlt">observed</span> spectra become steeper with increasing wind <span class="hlt">velocities</span> under quiet conditions, approach a spectral slope of -5/3 at a wind <span class="hlt">velocity</span> of 10 m s-1 and then roughly maintain this slope (-5/3) for even stronger winds. Our findings show an overall agreement with previous studies; furthermore, they provide a more complete climatology of frequency spectra of vertical wind <span class="hlt">velocities</span> under different wind conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830033826&hterms=DDD&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DDDD','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830033826&hterms=DDD&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DDDD"><span>Measurements of solar transition zone <span class="hlt">velocities</span> and line broadening using the ultraviolet spectrometer and polarimeter on the Solar Maximum Mission</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Simon, G.; Mein, P.; Vial, J. C.; Shine, R. A.; Woodgate, B. E.</p> <p>1982-01-01</p> <p>The UVSP instrument on SMM is able to <span class="hlt">observe</span> solar regions at two wavelengths in the same line with a band-pass of 0.3 A. Intensity and Doppler <span class="hlt">velocity</span> maps are derived. It is shown that the numerical values are sensitive to the adopted Doppler width and the range of <span class="hlt">velocities</span> is limited to within 30 km/sec. A method called Double Dopplergram Determination (DDD) is described for deriving both the Doppler width and the <span class="hlt">velocity</span> (up to 80 km/sec), and the main sources of uncertainties are discussed. To illustrate the method, a set of C IV 1548 A <span class="hlt">observations</span> is analyzed according to this procedure. The mean C IV Doppler width measured (0.15 A) is comparable to previous determinations. A relation is found between bright regions and down-flows. Large Doppler widths correspond to strong <span class="hlt">velocity</span> <span class="hlt">gradients</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015TCD.....9.4067A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015TCD.....9.4067A"><span><span class="hlt">Observations</span> of seasonal and diurnal glacier <span class="hlt">velocities</span> at Mount Rainier, Washington using terrestrial radar interferometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Allstadt, K. E.; Shean, D. E.; Campbell, A.; Fahnestock, M.; Malone, S. D.</p> <p>2015-07-01</p> <p>We present spatially continuous <span class="hlt">velocity</span> maps using repeat terrestrial radar interferometry (TRI) measurements to examine seasonal and diurnal dynamics of alpine glaciers at Mount Rainier, Washington. We show that the Nisqually and Emmons glaciers have small slope-parallel <span class="hlt">velocities</span> near the summit (< 0.2 m day-1), high <span class="hlt">velocities</span> over their upper and central regions (1.0-1.5 m day-1), and stagnant debris-covered regions near the terminus (< 0.05 m day-1). <span class="hlt">Velocity</span> uncertainties are as low as ±0.02-0.08 m day-1. We document a large seasonal <span class="hlt">velocity</span> decrease of 0.2-0.7 m day-1 (-25 to -50 %) from July to November for most of the Nisqually glacier, excluding the icefall, suggesting significant seasonal subglacial water storage under most of the glacier. We did not detect diurnal variability above the noise level. Preliminary 2-D ice flow modeling using TRI <span class="hlt">velocities</span> suggests that sliding accounts for roughly 91 and 99 % of the July <span class="hlt">velocity</span> field for the Emmons and Nisqually glaciers, respectively. We validate our <span class="hlt">observations</span> against recent in situ <span class="hlt">velocity</span> measurements and examine the long-term evolution of Nisqually glacier dynamics through comparisons with historical <span class="hlt">velocity</span> data. This study shows that repeat TRI measurements with > 10 km range can be used to investigate spatial and temporal variability of alpine glacier dynamics over large areas, including hazardous and inaccessible areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.H11F0905M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.H11F0905M"><span>A line source tracer test - a better method for assessing high groundwater <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Magal, E.; Weisbrod, N.; Yakirevich, A.; Kurtzman, D.; Yechieli, Y.</p> <p>2009-12-01</p> <p>A line source injection is suggested as an effective method for assessing groundwater <span class="hlt">velocities</span> and flow directions in subsurface characterized by high water fluxes. Modifying the common techniques of injecting a tracer into a well was necessary after frequently-used methods of natural and forced <span class="hlt">gradient</span> tracer tests ended with no reliable information on the local groundwater flow. In a field experiment, tracers were injected into 8-m long line injection system constructed below the water table almost perpendicular to the assumed flow direction. The injection system was divided to four separate segments (each 2 m long) enabling the injection of four different tracers along the line source. An array of five boreholes located in an area of 10x10 m downstream was used for monitoring the tracers' transport. Two dye tracers (Uranine and Na Naphthionate) were injected in a long pulse of several hours into two of the injection pipe segments and two tracers (Rhenium oxide and Gd-DTPA) were instantaneously injected to the other two segments. The tracers were detected 0.7 to 2.3 hours after injection in four of the five <span class="hlt">observation</span> wells, located 2.3 to 10 m from the injection system, respectively. Groundwater <span class="hlt">velocities</span> were calculated directly from the tracers' arrival times and by fitting the <span class="hlt">observed</span> breakthrough curves to simulations with one and two dimensions analytical solutions for conservative tracer transport. The groundwater <span class="hlt">velocity</span> was determined to be ~100 m/d. The longitudinal dispersivity value, generated from fitting the tracer breakthrough curves, was in a range of 0.2-3m. The groundwater flow direction was derived based on the arrival of the tracers and was found to be consistent with the apparent direction of the hydraulic <span class="hlt">gradient</span>. The hydraulic conductivity derived from the groundwater <span class="hlt">velocity</span> was ~1200 m/d, which is in the upper range of gravel sediment.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMEP51C0571B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMEP51C0571B"><span>Method to Rapidly Collect Thousands of <span class="hlt">Velocity</span> <span class="hlt">Observations</span> to Validate Million-Element 2D Hydrodynamic Models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barker, J. R.; Pasternack, G. B.; Bratovich, P.; Massa, D.; Reedy, G.; Johnson, T.</p> <p>2010-12-01</p> <p>Two-dimensional (depth-averaged) hydrodynamic models have existed for decades and are used to study a variety of hydrogeomorphic processes as well as to design river rehabilitation projects. Rapid computer and coding advances are revolutionizing the size and detail of 2D models. Meanwhile, advances in topo mapping and environmental informatics are providing the data inputs to drive large, detailed simulations. Million-element computational meshes are in hand. With simulations of this size and detail, the primary challenge has shifted to finding rapid and inexpensive means for testing model predictions against <span class="hlt">observations</span>. Standard methods for collecting <span class="hlt">velocity</span> data include boat-mounted ADCP and point-based sensors on boats or wading rods. These methods are labor intensive and often limited to a narrow flow range. Also, they generate small datasets at a few cross-sections, which is inadequate to characterize the statistical structure of the relation between predictions and <span class="hlt">observations</span>. Drawing on the long-standing oceanographic method of using drogues to track water currents, previous studies have demonstrated the potential of small dGPS units to obtain surface <span class="hlt">velocity</span> in rivers. However, dGPS is too inaccurate to test 2D models. Also, there is financial risk in losing drogues in rough currents. In this study, an RTK GPS unit was mounted onto a manned whitewater kayak. The boater positioned himself into the current and used floating debris to maintain a speed and heading consistent with the ambient surface flow field. RTK GPS measurements were taken ever 5 sec. From these positions, a 2D <span class="hlt">velocity</span> vector was obtained. The method was tested over ~20 km of the lower Yuba River in California in flows ranging from 500-5000 cfs, yielding 5816 <span class="hlt">observations</span>. To compare <span class="hlt">velocity</span> magnitude against the 2D model-predicted depth-averaged value, kayak-based surface values were scaled down by an optimized constant (0.72), which had no negative effect on regression analysis</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.466..892M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.466..892M"><span>SINFONI-HiZELS: the dynamics, merger rates and metallicity <span class="hlt">gradients</span> of 'typical' star-forming galaxies at z = 0.8-2.2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molina, J.; Ibar, Edo; Swinbank, A. M.; Sobral, D.; Best, P. N.; Smail, I.; Escala, A.; Cirasuolo, M.</p> <p>2017-04-01</p> <p>We present adaptive optics (AO) assisted SINFONI integral field unit (IFU) spectroscopy of 11 Hα emitting galaxies selected from the High-Z Emission Line Survey (HiZELS). We obtain spatially resolved dynamics on ˜kpc-scales of star-forming galaxies [stellar mass M⋆ = 109.5 - 10.5 M⊙ and star formation rate (SFR) = 2-30 M⊙ yr-1] near the peak of the cosmic star formation rate history. Combining these <span class="hlt">observations</span> with our previous SINFONI-HiZELS campaign, we construct a sample of 20 homogeneously selected galaxies with IFU AO-aided <span class="hlt">observations</span> - the 'SHiZELS' survey, with roughly equal number of galaxies per redshift slice, at z = 0.8, 1.47 and 2.23. We measure the dynamics and identify the major kinematic axis by modelling their <span class="hlt">velocity</span> fields to extract rotational curves and infer their inclination-corrected rotational <span class="hlt">velocities</span>. We explore the stellar mass Tully-Fisher relationship, finding that galaxies with higher <span class="hlt">velocity</span> dispersions tend to deviate from this relation. Using kinemetry analyses, we find that galaxy interactions might be the dominant mechanism controlling the star formation activity at z = 2.23 but they become gradually less important down to z = 0.8. Metallicity <span class="hlt">gradients</span> derived from the [N II]/Hα emission line ratio show a median negative <span class="hlt">gradient</span> for the SHiZELS survey of Δlog(O/H)/ΔR = -0.026 ± 0.008 dex kpc-1. We find that metal-rich galaxies tend to show negative <span class="hlt">gradients</span>, whereas metal-poor galaxies tend to exhibit positive metallicity <span class="hlt">gradients</span>. This result suggests that the accretion of pristine gas in the periphery of galaxies plays an important role in replenishing the gas in 'typical' star-forming galaxies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890001788','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890001788"><span>Migration arising from <span class="hlt">gradients</span> in shear stress: Particle distributions in Poiseuille flow</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Leighton, D. T., Jr.</p> <p>1988-01-01</p> <p>Experimental evidence for the existence of shear induced migration processes is reviewed and the mechanism by Leighton and Acrivos (1987b) is described in detail. The proposed mechanism is shown to lead to the existence of an additional shear induced migration in the presence of <span class="hlt">gradients</span> in shear stress such as would be found in Poiseuille flow, and which may be used to predict the amplitude of the <span class="hlt">observed</span> short-term viscosity increase. The concentration and <span class="hlt">velocity</span> profiles which result from such a migration are discussed in detail and are compared to the experimental <span class="hlt">observations</span> of Karnis, Goldsmith and Mason (1966).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...587A..97H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...587A..97H"><span>The Musca cloud: A 6 pc-long <span class="hlt">velocity</span>-coherent, sonic filament</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hacar, A.; Kainulainen, J.; Tafalla, M.; Beuther, H.; Alves, J.</p> <p>2016-03-01</p> <p>Filaments play a central role in the molecular clouds' evolution, but their internal dynamical properties remain poorly characterized. To further explore the physical state of these structures, we have investigated the kinematic properties of the Musca cloud. We have sampled the main axis of this filamentary cloud in 13CO and C18O (2-1) lines using APEX <span class="hlt">observations</span>. The different line profiles in Musca shows that this cloud presents a continuous and quiescent <span class="hlt">velocity</span> field along its ~6.5 pc of length. With an internal gas kinematics dominated by thermal motions (I.e. σNT/cs ≲ 1) and large-scale <span class="hlt">velocity</span> <span class="hlt">gradients</span>, these results reveal Musca as the longest <span class="hlt">velocity</span>-coherent, sonic-like object identified so far in the interstellar medium. The transonic properties of Musca present a clear departure from the predicted supersonic <span class="hlt">velocity</span> dispersions expected in the Larson's <span class="hlt">velocity</span> dispersion-size relationship, and constitute the first <span class="hlt">observational</span> evidence of a filament fully decoupled from the turbulent regime over multi-parsec scales. This publication is based on data acquired with the Atacama Pathfinder Experiment (APEX). APEX is a collaboration between the Max-Planck-Institut fuer Radioastronomie, the European Southern Observatory, and the Onsala Space Observatory (ESO programme 087.C-0583).The reduced datacubes as FITS files are only available at the CDS via anonymous ftp to http://cdsarc.u-strasbg.fr (ftp://130.79.128.5) or via http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/587/A97</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960045441','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960045441"><span>Effects of Spatial <span class="hlt">Gradients</span> on Electron Runaway Acceleration</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>MacNeice, Peter; Ljepojevic, N. N.</p> <p>1996-01-01</p> <p>The runaway process is known to accelerate electrons in many laboratory plasmas and has been suggested as an acceleration mechanism in some astrophysical plasmas, including solar flares. Current calculations of the electron <span class="hlt">velocity</span> distributions resulting from the runaway process are greatly restricted because they impose spatial homogeneity on the distribution. We have computed runaway distributions which include consistent development of spatial <span class="hlt">gradients</span> in the energetic tail. Our solution for the electron <span class="hlt">velocity</span> distribution is presented as a function of distance along a finite length acceleration region, and is compared with the equivalent distribution for the infinitely long homogenous system (i.e., no spatial <span class="hlt">gradients</span>), as considered in the existing literature. All these results are for the weak field regime. We also discuss the severe restrictiveness of this weak field assumption.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.G11A0470G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.G11A0470G"><span>Synthetic <span class="hlt">velocity</span> <span class="hlt">gradient</span> map of the San Francisco Bay region, California, supports use of average block <span class="hlt">velocities</span> to estimate fault slip rate where effective locking depth is small relative to inter-fault distance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Graymer, R. W.; Simpson, R. W.</p> <p>2014-12-01</p> <p>Graymer and Simpson (2013, AGU Fall Meeting) showed that in a simple 2D multi-fault system (vertical, parallel, strike-slip faults bounding blocks without strong material property contrasts) slip rate on block-bounding faults can be reasonably estimated by the difference between the mean <span class="hlt">velocity</span> of adjacent blocks if the ratio of the effective locking depth to the distance between the faults is 1/3 or less ("effective" locking depth is a synthetic parameter taking into account actual locking depth, fault creep, and material properties of the fault zone). To check the validity of that <span class="hlt">observation</span> for a more complex 3D fault system and a realistic distribution of <span class="hlt">observation</span> stations, we developed a synthetic suite of GPS <span class="hlt">velocities</span> from a dislocation model, with station location and fault parameters based on the San Francisco Bay region. Initial results show that if the effective locking depth is set at the base of the seismogenic zone (about 12-15 km), about 1/2 the interfault distance, the resulting synthetic <span class="hlt">velocity</span> <span class="hlt">observations</span>, when clustered, do a poor job of returning the input fault slip rates. However, if the apparent locking depth is set at 1/2 the distance to the base of the seismogenic zone, or about 1/4 the interfault distance, the synthetic <span class="hlt">velocity</span> field does a good job of returning the input slip rates except where the fault is in a strong restraining orientation relative to block motion or where block <span class="hlt">velocity</span> is not well defined (for example west of the northern San Andreas Fault where there are no <span class="hlt">observations</span> to the west in the ocean). The question remains as to where in the real world a low effective locking depth could usefully model fault behavior. Further tests are planned to define the conditions where average cluster-defined block <span class="hlt">velocities</span> can be used to reliably estimate slip rates on block-bounding faults. These rates are an important ingredient in earthquake hazard estimation, and another tool to provide them should be useful.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015TCry....9.2219A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015TCry....9.2219A"><span><span class="hlt">Observations</span> of seasonal and diurnal glacier <span class="hlt">velocities</span> at Mount Rainier, Washington, using terrestrial radar interferometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Allstadt, K. E.; Shean, D. E.; Campbell, A.; Fahnestock, M.; Malone, S. D.</p> <p>2015-12-01</p> <p>We present surface <span class="hlt">velocity</span> maps derived from repeat terrestrial radar interferometry (TRI) measurements and use these time series to examine seasonal and diurnal dynamics of alpine glaciers at Mount Rainier, Washington. We show that the Nisqually and Emmons glaciers have small slope-parallel <span class="hlt">velocities</span> near the summit (< 0.2 m day-1), high <span class="hlt">velocities</span> over their upper and central regions (1.0-1.5 m day-1), and stagnant debris-covered regions near the terminus (< 0.05 m day-1). <span class="hlt">Velocity</span> uncertainties are as low as ±0.02-0.08 m day-1. We document a large seasonal <span class="hlt">velocity</span> decrease of 0.2-0.7 m day-1 (-25 to -50 %) from July to November for most of the Nisqually Glacier, excluding the icefall, suggesting significant seasonal subglacial water storage under most of the glacier. We did not detect diurnal variability above the noise level. Simple 2-D ice flow modeling using TRI <span class="hlt">velocities</span> suggests that sliding accounts for 91 and 99 % of the July <span class="hlt">velocity</span> field for the Emmons and Nisqually glaciers with possible ranges of 60-97 and 93-99.5 %, respectively, when considering model uncertainty. We validate our <span class="hlt">observations</span> against recent in situ <span class="hlt">velocity</span> measurements and examine the long-term evolution of Nisqually Glacier dynamics through comparisons with historical <span class="hlt">velocity</span> data. This study shows that repeat TRI measurements with > 10 km range can be used to investigate spatial and temporal variability of alpine glacier dynamics over large areas, including hazardous and inaccessible areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..561..304H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..561..304H"><span><span class="hlt">Velocity</span> of water flow along saturated loess slopes under erosion effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Yuhan; Chen, Xiaoyan; Li, Fahu; Zhang, Jing; Lei, Tingwu; Li, Juan; Chen, Ping; Wang, Xuefeng</p> <p>2018-06-01</p> <p>Rainfall or snow-melted water recharge easily saturates loose top soils with a less permeable underlayer, such as cultivated soil slope and partially thawed top soil layer, and thus, may influence the <span class="hlt">velocity</span> of water flow. This study suggested a methodology and device system to supply water from the bottom soil layer at the different locations of slopes. Water seeps into and saturates the soil, when the water level is controlled at the same height of the soil surface. The structures and functions of the device, the components, and the operational principles are described in detail. A series of laboratory experiments were conducted under slope <span class="hlt">gradients</span> of 5°, 10°, 15°, and 20° and flow rates of 2, 4, and 8 L min-1 to measure the water flow <span class="hlt">velocities</span> over eroding and non-eroded loess soil slopes, under saturated conditions by using electrolyte tracing. Results showed that flow <span class="hlt">velocities</span> on saturated slopes were 17% to 88% greater than those on non-saturated slopes. Flow <span class="hlt">velocity</span> increased rapidly under high flow rates and slope <span class="hlt">gradients</span>. Saturation conditions were suitable in maintaining smooth rill geomorphology and causing fast water flow. The saturated soil slope had a lubricant effect on the soil surface to reduce the frictional force, resulting in high flow <span class="hlt">velocity</span>. The flow <span class="hlt">velocities</span> of eroding rills under different slope <span class="hlt">gradients</span> and flow rates were approximately 14% to 33% lower than those of non-eroded rills on saturated loess slopes. Compared with that on a saturated loess slope, the eroding rill on a non-saturated loess slope can produce headcuts to reduce the flow <span class="hlt">velocity</span>. This study helps understand the hydrodynamics of soil erosion and sediment transportation of saturated soil slopes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApPhL.107k1603P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApPhL.107k1603P"><span><span class="hlt">Gradient</span> induced liquid motion on laser structured black Si surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paradisanos, I.; Fotakis, C.; Anastasiadis, S. H.; Stratakis, E.</p> <p>2015-09-01</p> <p>This letter reports on the femtosecond laser fabrication of <span class="hlt">gradient</span>-wettability micro/nano-patterns on Si surfaces. The dynamics of directional droplet spreading on the surface tension <span class="hlt">gradients</span> developed is systematically investigated and discussed. It is shown that microdroplets on the patterned surfaces spread at a maximum speed of 505 mm/s, which is the highest <span class="hlt">velocity</span> demonstrated so far for liquid spreading on a surface tension <span class="hlt">gradient</span> in ambient conditions. The application of the proposed laser patterning technique for the precise fabrication of surface tension <span class="hlt">gradients</span> for open microfluidic systems, liquid management in fuel cells, and drug delivery is envisaged.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012DSRI...69...36U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012DSRI...69...36U"><span>The hydrography of the Mozambique Channel from six years of continuous temperature, salinity, and <span class="hlt">velocity</span> <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ullgren, J. E.; van Aken, H. M.; Ridderinkhof, H.; de Ruijter, W. P. M.</p> <p>2012-11-01</p> <p>Temperature, salinity and <span class="hlt">velocity</span> data are presented, along with the estimated volume transport, from seven full-length deep sea moorings placed across the narrowest part of the Mozambique Channel, southwest Indian Ocean, during the period November 2003 to December 2009. The dominant water mass in the upper layer is Sub-Tropical Surface Water (STSW) which overlies South Indian Central Water (SICW), and is normally capped by fresher Tropical Surface Water (TSW). Upper ocean salinity increased through 2005 as a result of saline STSW taking up a relatively larger part of the upper layer, at the expense of TSW. Upper waters are on average warmer and lighter in the central Channel than on the sides. Throughout the upper 1.5 km of the water column there is large hydrographic variability, short-term as well as interannual, and in particular at frequencies (four to seven cycles per year) associated with the southward passage of anticyclonic Mozambique Channel eddies. The eddies have a strong T-S signal, in the upper and central waters as well as on the intermediate level, as the eddies usually carry saline Red Sea Water (RSW) in their core. While the interannual frequency band displays an east-west <span class="hlt">gradient</span> with higher temperature variance on the western side, the eddy frequency band shows highest variance in the centre of the Channel, where the eddy band contains about 40% of the total isopycnal hydrographic variability. Throughout the >6 years of measurements, the frequency and characteristics of eddies vary between periods, both in terms of strength and vertical structure of eddy T-S signals. These changes contribute to the interannual variability of water mass properties: an increase in central water salinity to a maximum in late 2007 coincided with a period of unusually frequent eddies with strong salinity signals. The warmest and most saline deep water is found within the northward flowing Mozambique Undercurrent, on the western side of the Channel. The Undercurrent</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhPl...19e2303Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhPl...19e2303Q"><span>Tripolar vortex formation in dense quantum plasma with ion-temperature-<span class="hlt">gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qamar, Anisa; Ata-ur-Rahman, Mirza, Arshad M.</p> <p>2012-05-01</p> <p>We have derived system of nonlinear equations governing the dynamics of low-frequency electrostatic toroidal ion-temperature-<span class="hlt">gradient</span> mode for dense quantum magnetoplasma. For some specific profiles of the equilibrium density, temperature, and ion <span class="hlt">velocity</span> <span class="hlt">gradients</span>, the nonlinear equations admit a stationary solution in the form of a tripolar vortex. These results are relevant to understand nonlinear structure formation in dense quantum plasmas in the presence of equilibrium ion-temperature and density <span class="hlt">gradients</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880013717','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880013717"><span>A statistical investigation of the single-point pdf of <span class="hlt">velocity</span> and vorticity based on direct numerical simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mortazavi, M.; Kollmann, W.; Squires, K.</p> <p>1987-01-01</p> <p>Vorticity plays a fundamental role in turbulent flows. The dynamics of vorticity in turbulent flows and the effect on single-point closure models were investigated. The approach was to use direct numerical simulations of turbulent flows to investigate the pdf of <span class="hlt">velocity</span> and vorticity. The preliminary study of homogeneous shear flow has shown that the expectation of the fluctuating pressure <span class="hlt">gradient</span>, conditioned with a <span class="hlt">velocity</span> component, is linear in the <span class="hlt">velocity</span> component, and that the coefficient is independent of <span class="hlt">velocity</span> and vorticity. In addition, the work shows that the expectation of the pressure <span class="hlt">gradient</span>, conditioned with a vorticity component, is essentially zero.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22034696-bulge-radial-velocity-assay-brava-ii-complete-sample-data-release','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22034696-bulge-radial-velocity-assay-brava-ii-complete-sample-data-release"><span>THE BULGE RADIAL <span class="hlt">VELOCITY</span> ASSAY (BRAVA). II. COMPLETE SAMPLE AND DATA RELEASE</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kunder, Andrea; De Propris, Roberto; Stubbs, Scott A.</p> <p>2012-03-15</p> <p>We present new radial <span class="hlt">velocity</span> measurements from the Bulge Radial <span class="hlt">Velocity</span> Assay, a large-scale spectroscopic survey of M-type giants in the Galactic bulge/bar region. The sample of {approx}4500 new radial <span class="hlt">velocities</span>, mostly in the region -10 Degree-Sign < l < +10 Degree-Sign and b Almost-Equal-To -6 Degree-Sign , more than doubles the existent published data set. Our new data extend our rotation curve and <span class="hlt">velocity</span> dispersion profile to +20 Degree-Sign , which is {approx}2.8 kpc from the Galactic center. The new data confirm the cylindrical rotation <span class="hlt">observed</span> at -6 Degree-Sign and -8 Degree-Sign and are an excellent fit to themore » Shen et al. N-body bar model. We measure the strength of the TiO{epsilon} molecular band as a first step toward a metallicity ranking of the stellar sample, from which we confirm the presence of a vertical abundance <span class="hlt">gradient</span>. Our survey finds no strong evidence of previously unknown kinematic streams. We also publish our complete catalog of radial <span class="hlt">velocities</span>, photometry, TiO band strengths, and spectra, which is available at the Infrared Science Archive as well as at UCLA.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770024497','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770024497"><span>Representation of turbulent shear stress by a product of mean <span class="hlt">velocity</span> differences</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Braun, W. H.</p> <p>1977-01-01</p> <p>A quadratic form in the mean <span class="hlt">velocity</span> for the turbulent shear stress is presented. It is expressed as the product of two <span class="hlt">velocity</span> differences whose roots are the maximum <span class="hlt">velocity</span> in the flow and a cutoff <span class="hlt">velocity</span> below which the turbulent shear stress vanishes. Application to pipe and channel flows yields the centerline <span class="hlt">velocity</span> as a function of pressure <span class="hlt">gradient</span>, as well as the <span class="hlt">velocity</span> profile. The flat plate, boundary-layer problem is solved by a system of integral equations to obtain friction coefficient, displacement thickness, and momentum-loss thickness. Comparisons are made with experiment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880062503&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dvertical%2Bheight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880062503&hterms=vertical+height&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dvertical%2Bheight"><span><span class="hlt">Observations</span> of vertical <span class="hlt">velocities</span> in the tropical upper troposphere and lower stratosphere using the Arecibo 430-MHz radar</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cornish, C. R.</p> <p>1988-01-01</p> <p>The first clear-air <span class="hlt">observations</span> of vertical <span class="hlt">velocities</span> in the tropical upper troposphere and lower stratosphere (8-22 km) using the Arecibo 430-MHz radar are presented. Oscillations in the vertical <span class="hlt">velocity</span> near the Brunt-Vaisala period are <span class="hlt">observed</span> in the lower stratosphere during the 12-hour <span class="hlt">observation</span> period. Frequency power spectra from the vertical <span class="hlt">velocity</span> time series show a slope between -0.5 and -1.0. Vertical wave number spectra computed from the height profiles of vertical <span class="hlt">velocities</span> have slopes between -1.0 and -1.5. These <span class="hlt">observed</span> slopes do not agree well with the slopes of +1/3 and -2.5 for frequency and vertical wave number spectra, respectively, predicted by a universal gravity-wave spectrum model. The spectral power of wave number spectra of a radial beam directed 15 deg off-zenith is enhanced by an order of magnitude over the spectral power levels of the vertical beam. This enhancement suggests that other geophysical processes besides gravity waves are present in the horizontal flow. The steepening of the wave number spectrum of the off-vertical beam in the lower stratosphere to near -2.0 is attributed to a quasi-inertial period wave, which was present in the horizontal flow during the <span class="hlt">observation</span> period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555178','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555178"><span>Motion Driven by Strain <span class="hlt">Gradient</span> Fields</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Chao; Chen, Shaohua</p> <p>2015-01-01</p> <p>A new driving mechanism for direction-controlled motion of nano-scale objects is proposed, based on a model of stretching a graphene strip linked to a rigid base with linear springs of identical stiffness. We find that the potential energy difference induced by the strain <span class="hlt">gradient</span> field in the graphene strip substrate can generate sufficient force to overcome the static and kinetic friction forces between the nano-flake and the strip substrate, resulting in the nanoscale flake motion in the direction of <span class="hlt">gradient</span> reduction. The dynamics of the nano-flake can be manipulated by tuning the stiffness of linear springs, stretching <span class="hlt">velocity</span> and the flake size. This fundamental law of directional motion induced by strain <span class="hlt">gradient</span> could be very useful for promising designs of nanoscale manipulation, transportation and smart surfaces. PMID:26323603</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvD..97f3013Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvD..97f3013Z"><span>Constraint on the <span class="hlt">velocity</span> dependent dark matter annihilation cross section from gamma-ray and kinematic <span class="hlt">observations</span> of ultrafaint dwarf galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Yi; Bi, Xiao-Jun; Yin, Peng-Fei; Zhang, Xinmin</p> <p>2018-03-01</p> <p>Searching for γ rays from dwarf spheroidal galaxies (dSphs) is a promising approach to detect dark matter (DM) due to the high DM densities and low baryon components in dSphs. The Fermi-LAT <span class="hlt">observations</span> from dSphs have set stringent constraints on the <span class="hlt">velocity</span> independent annihilation cross section. However, the constraints from dSphs may change in <span class="hlt">velocity</span> dependent annihilation scenarios because of the different <span class="hlt">velocity</span> dispersions in galaxies. In this work, we study how to set constraints on the <span class="hlt">velocity</span> dependent annihilation cross section from the combined Fermi-LAT <span class="hlt">observations</span> of dSphs with the kinematic data. In order to calculate the γ ray flux from the dSph, the correlation between the DM density profile and <span class="hlt">velocity</span> dispersion at each position should be taken into account. We study such correlation and the relevant uncertainty from kinematic <span class="hlt">observations</span> by performing a Jeans analysis. Using the <span class="hlt">observational</span> results of three ultrafaint dSphs with large J-factors, including Willman 1, Reticulum II, and Triangulum II, we set constraints on the p-wave annihilation cross section in the Galaxy as an example.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeoJI.207.1493K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeoJI.207.1493K"><span>Two-receiver measurements of phase <span class="hlt">velocity</span>: cross-validation of ambient-noise and earthquake-based <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kästle, Emanuel D.; Soomro, Riaz; Weemstra, Cornelis; Boschi, Lapo; Meier, Thomas</p> <p>2016-12-01</p> <p>Phase <span class="hlt">velocities</span> derived from ambient-noise cross-correlation are compared with phase <span class="hlt">velocities</span> calculated from cross-correlations of waveform recordings of teleseismic earthquakes whose epicentres are approximately on the station-station great circle. The comparison is conducted both for Rayleigh and Love waves using over 1000 station pairs in central Europe. We describe in detail our signal-processing method which allows for automated processing of large amounts of data. Ambient-noise data are collected in the 5-80 s period range, whereas teleseismic data are available between about 8 and 250 s, resulting in a broad common period range between 8 and 80 s. At intermediate periods around 30 s and for shorter interstation distances, phase <span class="hlt">velocities</span> measured from ambient noise are on average between 0.5 per cent and 1.5 per cent lower than those <span class="hlt">observed</span> via the earthquake-based method. This discrepancy is small compared to typical phase-<span class="hlt">velocity</span> heterogeneities (10 per cent peak-to-peak or more) <span class="hlt">observed</span> in this period range.We nevertheless conduct a suite of synthetic tests to evaluate whether known biases in ambient-noise cross-correlation measurements could account for this discrepancy; we specifically evaluate the effects of heterogeneities in source distribution, of azimuthal anisotropy in surface-wave <span class="hlt">velocity</span> and of the presence of near-field, rather than far-field only, sources of seismic noise. We find that these effects can be quite important comparing individual station pairs. The systematic discrepancy is presumably due to a combination of factors, related to differences in sensitivity of earthquake versus noise data to lateral heterogeneity. The data sets from both methods are used to create some preliminary tomographic maps that are characterized by <span class="hlt">velocity</span> heterogeneities of similar amplitude and pattern, confirming the overall agreement between the two measurement methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790039195&hterms=oso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doso','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790039195&hterms=oso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doso"><span>The height variation of supergranular <span class="hlt">velocity</span> fields determined from simultaneous OSO 8 satellite and ground-based <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>November, L. J.; Toomre, J.; Gebbie, K. B.; Simon, G. W.</p> <p>1979-01-01</p> <p>Results are reported for simultaneous satellite and ground-based <span class="hlt">observations</span> of supergranular <span class="hlt">velocities</span> in the sun, which were made using a UV spectrometer aboard OSO 8 and a diode-array instrument operating at the exit slit of an echelle spectrograph attached to a vacuum tower telescope. <span class="hlt">Observations</span> of the steady Doppler <span class="hlt">velocities</span> seen toward the limb in the middle chromosphere and the photosphere are compared; the <span class="hlt">observed</span> spectral lines of Si II at 1817 A and Fe I at 5576 A are found to differ in height of formation by about 1400 km. The results show that supergranular motions are able to penetrate at least 11 density scale heights into the middle chromosphere, that the patterns of motion correlate well with the cellular structure seen in the photosphere, and that the motion increases from about 800 m/s in the photosphere to at least 3000 m/s in the middle chromosphere. These <span class="hlt">observations</span> imply that supergranular <span class="hlt">velocities</span> should be evident in the transition region and that strong horizontal shear layers in supergranulation should produce turbulence and internal gravity waves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.475.2697P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.475.2697P"><span>Revisiting the stellar <span class="hlt">velocity</span> ellipsoid-Hubble-type relation: <span class="hlt">observations</span> versus simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pinna, F.; Falcón-Barroso, J.; Martig, M.; Martínez-Valpuesta, I.; Méndez-Abreu, J.; van de Ven, G.; Leaman, R.; Lyubenova, M.</p> <p>2018-04-01</p> <p>The stellar <span class="hlt">velocity</span> ellipsoid (SVE) in galaxies can provide important information on the processes that participate in the dynamical heating of their disc components (e.g. giant molecular clouds, mergers, spiral density waves, and bars). Earlier findings suggested a strong relation between the shape of the disc SVE and Hubble type, with later-type galaxies displaying more anisotropic ellipsoids and early types being more isotropic. In this paper, we revisit the strength of this relation using an exhaustive compilation of <span class="hlt">observational</span> results from the literature on this issue. We find no clear correlation between the shape of the disc SVE and morphological type, and show that galaxies with the same Hubble type display a wide range of vertical-to-radial <span class="hlt">velocity</span> dispersion ratios. The points are distributed around a mean value and scatter of σz/σR = 0.7 ± 0.2. With the aid of numerical simulations, we argue that different mechanisms might influence the shape of the SVE in the same manner and that the same process (e.g. mergers) does not have the same impact in all the galaxies. The complexity of the <span class="hlt">observational</span> picture is confirmed by these simulations, which suggest that the vertical-to-radial axis ratio of the SVE is not a good indicator of the main source of disc heating. Our analysis of those simulations also indicates that the <span class="hlt">observed</span> shape of the disc SVE may be affected by several processes simultaneously and that the signatures of some of them (e.g. mergers) fade over time.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1911495B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1911495B"><span>Control of topography <span class="hlt">gradients</span> on residence time distributions, mixing dynamics and reactive hotspot development</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bandopadhyay, Aditya; Le Borgne, Tanguy; Davy, Philippe</p> <p>2017-04-01</p> <p>Topography-driven subsurface flows are thought to play a central role in determining solute turnover and biogeochemical processes at different scales in the critical zone, including river-hyporheic zone exchanges, hillslope solute transport and reactions, and catchment biogeochemical cycles. Hydraulic head <span class="hlt">gradients</span>, induced by topography <span class="hlt">gradients</span> at different scales, generate a distribution of streamlines at depth, dictating the spatial distribution of redox sensitive species, the magnitude of surface water - ground water exchanges and ultimately the source/sink function of the subsurface. Flow <span class="hlt">velocities</span> generally decrease with depth, leading to broad residence time distributions, which have been shown to affect river chemistry and geochemical reactions in catchments. In this presentation, we discuss the impact of topography-driven flows on mixing processes and the formation of localized reactive hotspots. For this, we solve analytically the coupled flow, mixing and reaction equations in two-dimensional vertical cross-sections of subsurface domains with different topography <span class="hlt">gradients</span>. For a given topography <span class="hlt">gradient</span>, we derive the spatial distribution of subsurface <span class="hlt">velocities</span>, the rates of solute mixing accross streamlines and the induced kinetics of redox, precipitation and dissolution reactions using a Lagrangian approach (Le Borgne et al. 2014). We demonstrate that vertical <span class="hlt">velocity</span> profiles driven by topography variations, act effectively as shear flows, hence stretching continuously the mixing fronts between recently infiltrated and resident water (Bandopadhyay et al. 2017). We thus derive analytical expressions for residence time distributions, mixing rates and kinetics of chemical reactions as a function of the topography <span class="hlt">gradients</span>. We show that the rates dissolution and precipitation reactions are significantly enhanced by the existence of vertical <span class="hlt">velocity</span> <span class="hlt">gradients</span> and that reaction rates reach a maximum in a localized subsurface reactive layer, whose</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmpL..95G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmpL..95G"><span>Revealing the <span class="hlt">velocity</span> structure of the filamentary nebula in NGC 1275 in its entirety</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gendron-Marsolais, M.; Hlavacek-Larrondo, J.; Martin, T. B.; Drissen, L.; McDonald, M.; Fabian, A. C.; Edge, A. C.; Hamer, S. L.; McNamara, B.; Morrison, G.</p> <p>2018-05-01</p> <p>We have produced for the first time a detailed <span class="hlt">velocity</span> map of the giant filamentary nebula surrounding NGC 1275, the Perseus cluster's brightest galaxy, and revealed a previously unknown rich <span class="hlt">velocity</span> structure across the entire nebula. These new <span class="hlt">observations</span> were obtained with the optical imaging Fourier transform spectrometer SITELLE at CFHT. With its wide field of view (˜11'×11'), SITELLE is the only integral field unit spectroscopy instrument able to cover the 80 kpc×55 kpc (3.8'×2.6') large nebula in NGC 1275. Our analysis of these <span class="hlt">observations</span> shows a smooth radial <span class="hlt">gradient</span> of the [N II]λ6583/Hα line ratio, suggesting a change in the ionization mechanism and source across the nebula. The <span class="hlt">velocity</span> map shows no visible general trend or rotation, indicating that filaments are not falling uniformly onto the galaxy, nor being uniformly pulled out from it. Comparison between the physical properties of the filaments and Hitomi measurements of the X-ray gas dynamics in Perseus are also explored.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhBio..15b6010B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhBio..15b6010B"><span>Bidirectional transport model of morphogen <span class="hlt">gradient</span> formation via cytonemes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bressloff, Paul C.; Kim, Hyunjoong</p> <p>2018-03-01</p> <p>Morphogen protein <span class="hlt">gradients</span> play an important role in the spatial regulation of patterning during embryonic development. The most commonly accepted mechanism for <span class="hlt">gradient</span> formation is diffusion from a source combined with degradation. Recently, there has been growing interest in an alternative mechanism, which is based on the direct delivery of morphogens along thin, actin-rich cellular extensions known as cytonemes. In this paper, we develop a bidirectional motor transport model for the flux of morphogens along cytonemes, linking a source cell to a one-dimensional array of target cells. By solving the steady-state transport equations, we show how a morphogen <span class="hlt">gradient</span> can be established, and explore how the mean <span class="hlt">velocity</span> of the motors affects properties of the morphogen <span class="hlt">gradient</span> such as accumulation time and robustness. In particular, our analysis suggests that in order to achieve robustness with respect to changes in the rate of synthesis of morphogen, the mean <span class="hlt">velocity</span> has to be negative, that is, retrograde flow or treadmilling dominates. Thus the potential targeting precision of cytonemes comes at an energy cost. We then study the effects of non-uniformly allocating morphogens to the various cytonemes projecting from a source cell. This competition for resources provides a potential regulatory control mechanism not available in diffusion-based models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160011108','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160011108"><span>Evolution of a Planar Wake in Adverse Pressure <span class="hlt">Gradient</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Driver, David M.; Mateer, George G.</p> <p>2016-01-01</p> <p>In the interest of improving the predictability of high-lift systems at maximum lift conditions, a series of fundamental experiments were conducted to study the effects of adverse pressure <span class="hlt">gradient</span> on a wake flow. Mean and fluctuating <span class="hlt">velocities</span> were measured with a two-component laser-Doppler velocimeter. Data were obtained for several cases of adverse pressure <span class="hlt">gradient</span>, producing flows ranging from no reversed flow to massively reversed flow. While the turbulent Reynolds stresses increase with increasing size of the reversed flow region, the <span class="hlt">gradient</span> of Reynolds stress does not. Computations using various turbulence models were unable to reproduce the reversed flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1893c0133O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1893c0133O"><span>Numerical modeling of the destruction of steel plates with a <span class="hlt">gradient</span> substrate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Orlov, M. Yu.; Glazyrin, V. P.; Orlov, Yu. N.</p> <p>2017-10-01</p> <p>The paper presents the results of numerical simulation of the shock loading process of steel barriers with a <span class="hlt">gradient</span> substrate. In an elastic plastic axisymmetric statement, a shock is simulated along the normal in the range of initial <span class="hlt">velocities</span> up to 300 m / s. A range of initial <span class="hlt">velocities</span> was revealed, in which the presence of a substrate "saved" the obstacle from spallation. New tasks were announced to deepen scientific knowledge about the behavior of unidirectional <span class="hlt">gradient</span> barriers at impact. The results of calculations are obtained in the form of graphs, calculated configurations of the "impact - barrier" and tables.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890046357&hterms=1087&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3D%2526%25231087','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890046357&hterms=1087&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3D%2526%25231087"><span>Acoustic waves in gases with strong pressure <span class="hlt">gradients</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zorumski, William E.</p> <p>1989-01-01</p> <p>The effect of strong pressure <span class="hlt">gradients</span> on the acoustic modes (standing waves) of a rectangular cavity is investigated analytically. When the cavity response is represented by a sum of modes, each mode is found to have two resonant frequencies. The lower frequency is near the Viaesaela-Brundt frequency, which characterizes the buoyant effect, and the higher frequency is above the ordinary acoustic resonance frequency. This finding shows that the propagation <span class="hlt">velocity</span> of the acoustic waves is increased due to the pressure <span class="hlt">gradient</span> effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMDI11C2613C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMDI11C2613C"><span>Seismic <span class="hlt">Observations</span> of the Mid-Pacific Large Low Shear <span class="hlt">Velocity</span> Province</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chan, A.; Helmberger, D. V.; Sun, D.; Li, D.; Jackson, J. M.</p> <p>2015-12-01</p> <p>Seismic data from earthquakes originating in the Fiji-Tonga region exhibits waveform complexity of a number of phases which may be attributed to various structures along ray paths to stations of USArray, including anomalous structures at the core-mantle boundary. The data shows variation in multipathing, that is, the presence of secondary arrivals following the S phase at diffracted distances (Sdiff) which suggests that the waveform complexity is due to structures at the eastern edge of the mid-Pacific Large Low Shear <span class="hlt">Velocity</span> Province (LLSVP). This study examines data from earthquake events while the Transportable Array portion of USArray was situated in the midwest United States, reinforcing previous studies that indicate late arrivals occurring as long as 26 seconds after the primary arrivals (To et al., 2011). Using earth flattening transformations and finite difference methods, simulations of tapered wedge structures of low <span class="hlt">velocity</span> material allow for wave energy trapping, producing the <span class="hlt">observed</span> waveform complexity and delayed arrivals at large distances, with such structures having characteristic properties of, for example, a height of 70 km, in-plane extent more than 1000 km, and shear wave <span class="hlt">velocity</span> drop of 3% at the top to 15% at the bottom relative to PREM. Differential arrival times for SH and SV components suggest anisotropy and possible wave propagation through downgoing slabs beneath the source region. The arrivals of the SPdKS phase further support the presence of an ultra-low <span class="hlt">velocity</span> zone (ULVZ) within a two-humped LLSVP. Some systematic delays in arrival times of multiple phases for distances less than 102º are accounted for and attributed to the presence of a mantle slab underneath the continental United States. Comparisons to seismic data from earthquakes originating from other locations further constrain depths of the deep mantle structures. Possible explanations include iron-enrichment of deep mantle phases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990hst..prop.2238B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990hst..prop.2238B"><span>Lyman-Alpha <span class="hlt">Observations</span> of High Radial <span class="hlt">Velocity</span> Stars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bookbinder, Jay</p> <p>1990-12-01</p> <p>H I LYMAN -ALPHA (LY-A) IS ONE OF THE MOST IMPORTANT LINES EMITTED BY PLASMA IN THE TEMPERATURE RANGE OF 7000 TO 10 TO THE FIFTH POWER K IN LATE-TYPE STARS. IT IS A MAJOR COMPONENT OF THE TOTAL RADIATIVE LOSS RATE, AND IT PLAYS A CRUCIAL ROLE IN DETERMINING THE ATMOSPHERIC STRUCTURE AND IN FLUORESCING OTHER UV LINES. YET IT IS ALSO THE LEAST STUDIED MAJOR LINE IN THE FAR UV, BECAUSE MOST OF THE LINE FLUX IS ABSORBED BY THE ISM ALONG THE LINE OF SIGHT AND BECAUSE IT IS STRONGLY COMTAMINATED BY THE GEOCORONAL BACKGROUND. A KNOWLEDGE OF THE Ly-A PROFILE IS ALSO IMPORTANT FOR STUDIES OF DEUTERIUM IN THE INTERSTELLAR MEDIUM. BY <span class="hlt">OBSERVING</span> HIGH RADIAL <span class="hlt">VELOCITY</span> STARS WE WILL OBTAIN FOR THE FIRST TIME HIGH RESOLUTION SPECTRA OF THE CORE OF A STELLAR H I LYMAN-A EMISSION LINE PROFILE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900040580&hterms=competence&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcompetence','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900040580&hterms=competence&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcompetence"><span>Angular <span class="hlt">velocity</span> discrimination</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kaiser, Mary K.</p> <p>1990-01-01</p> <p>Three experiments designed to investigate the ability of naive <span class="hlt">observers</span> to discriminate rotational <span class="hlt">velocities</span> of two simultaneously viewed objects are described. Rotations are constrained to occur about the x and y axes, resulting in linear two-dimensional image trajectories. The results indicate that <span class="hlt">observers</span> can discriminate angular <span class="hlt">velocities</span> with a competence near that for linear <span class="hlt">velocities</span>. However, perceived angular rate is influenced by structural aspects of the stimuli.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RaSc...52..951S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RaSc...52..951S"><span>Evaluation of extreme ionospheric total electron content <span class="hlt">gradient</span> associated with plasma bubbles for GNSS Ground-Based Augmentation System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saito, S.; Yoshihara, T.</p> <p>2017-08-01</p> <p>Associated with plasma bubbles, extreme spatial <span class="hlt">gradients</span> in ionospheric total electron content (TEC) were <span class="hlt">observed</span> on 8 April 2008 at Ishigaki (24.3°N, 124.2°E, +19.6° magnetic latitude), Japan. The largest <span class="hlt">gradient</span> was 3.38 TECU km-1 (total electron content unit, 1 TECU = 1016 el m-2), which is equivalent to an ionospheric delay <span class="hlt">gradient</span> of 540 mm km-1 at the GPS L1 frequency (1.57542 GHz). This value is confirmed by using multiple estimating methods. The <span class="hlt">observed</span> value exceeds the maximum ionospheric <span class="hlt">gradient</span> that has ever been <span class="hlt">observed</span> (412 mm km-1 or 2.59 TECU km-1) to be associated with a severe magnetic storm. It also exceeds the assumed maximum value (500 mm km-1 or 3.08 TECU km-1) which was used to validate the draft international standard for Global Navigation Satellite System (GNSS) Ground-Based Augmentation Systems (GBAS) to support Category II/III approaches and landings. The steepest part of this extreme <span class="hlt">gradient</span> had a scale size of 5.3 km, and the front-normal <span class="hlt">velocities</span> were estimated to be 71 m s-1 with a wavefront-normal direction of east-northeastward. The total width of the transition region from outside to inside the plasma bubble was estimated to be 35.3 km. The <span class="hlt">gradient</span> of relatively small spatial scale size may fall between an aircraft and a GBAS ground subsystem and may be undetectable by both aircraft and ground.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005ApJ...618..679H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005ApJ...618..679H"><span>Chandra <span class="hlt">Observations</span> of Low <span class="hlt">Velocity</span> Dispersion Groups</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Helsdon, Stephen F.; Ponman, Trevor J.; Mulchaey, J. S.</p> <p>2005-01-01</p> <p>Deviations of galaxy groups from cluster scaling relations can be understood in terms of an excess of entropy in groups. The main effect of this excess is to reduce the density and thus the luminosity of the intragroup gas. Given this, groups should also show a steep relationship between X-ray luminosity and <span class="hlt">velocity</span> dispersion. However, previous work suggests that this is not the case, with many measuring slopes flatter than the cluster relation. Examining the group LX-σ relation shows that much of the flattening is caused by a small subset of groups that show very high X-ray luminosities for their <span class="hlt">velocity</span> dispersions (or vice versa). Detailed Chandra study of two such groups shows that earlier ROSAT results were subject to significant (~30%-40%) point-source contamination but confirm that a significant hot intergalactic medium is present in these groups, although these are two of the coolest systems in which intergalactic X-ray emission has been detected. Their X-ray properties are shown to be broadly consistent with those of other galaxy groups, although the gas entropy in NGC 1587 is unusually low, and its X-ray luminosity is correspondingly high for its temperature when compared with most groups. This leads us to suggest that the <span class="hlt">velocity</span> dispersion in these systems has been reduced in some way, and we consider how this might have come about.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960015858','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960015858"><span>A Study of the Development of Steady and Periodic Unsteady Turbulent Wakes Through Curved Channels at Positive, Zero, and Negative Streamwise Pressure <span class="hlt">Gradients</span>, Part 1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schobeiri, M. T.; John, J.</p> <p>1996-01-01</p> <p>The turbomachinery wake flow development is largely influenced by streamline curvature and streamwise pressure <span class="hlt">gradient</span>. The objective of this investigation is to study the development of the wake under the influence of streamline curvature and streamwise pressure <span class="hlt">gradient</span>. The experimental investigation is carried out in two phases. The first phase involves the study of the wake behind a stationary circular cylinder (steady wake) in curved channels at positive, zero, and negative streamwise pressure <span class="hlt">gradients</span>. The mean <span class="hlt">velocity</span> and Reynolds stress components are measured using a X-hot-film probe. The measured quantities obtained in probe coordinates are transformed to a curvilinear coordinate system along the wake centerline and are presented in similarity coordinates. The results of the steady wakes suggest strong asymmetry in <span class="hlt">velocity</span> and Reynolds stress components. However, the <span class="hlt">velocity</span> defect profiles in similarity coordinates are almost symmetrical and follow the same distribution as the zero pressure <span class="hlt">gradient</span> straight wake. The results of Reynolds stress distributions show higher values on the inner side of the wake than the outer side. Other quantities, including the decay of maximum <span class="hlt">velocity</span> defect, growth of wake width, and wake integral parameters, are also presented for the three different pressure <span class="hlt">gradient</span> cases of steady wake. The decay rate of <span class="hlt">velocity</span> defect is fastest for the negative streamwise pressure <span class="hlt">gradient</span> case and slowest for the positive pressure <span class="hlt">gradient</span> case. Conversely, the growth of the wake width is fastest for the positive streamwise pressure <span class="hlt">gradient</span> case and slowest for the negative streamwise pressure <span class="hlt">gradient</span>. The second phase studies the development of periodic unsteady wakes generated by the circular cylinders of the rotating wake generator in a curved channel at zero streamwise pressure <span class="hlt">gradient</span>. Instantaneous <span class="hlt">velocity</span> components of the periodic unsteady wakes, measured with a stationary X-hot-film probe, are analyzed by the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMSA31B4097M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSA31B4097M"><span>Comparison with the horizontal phase <span class="hlt">velocity</span> distribution of gravity waves <span class="hlt">observed</span> airglow imaging data of different sampling periods</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Matsuda, T. S.; Nakamura, T.; Ejiri, M. K.; Tsutsumi, M.; Shiokawa, K.</p> <p>2014-12-01</p> <p>Atmospheric gravity waves (AGWs), which are generated in the lower atmosphere, transport significant amount of energy and momentum into the mesosphere and lower thermosphere. Among many parameters to characterize AGWs, horizontal phase <span class="hlt">velocity</span> is very important to discuss the vertical propagation. Airglow imaging is a useful technique for investigating the horizontal structures of AGWs around mesopause. There are many airglow imagers operated all over the world, and a large amount of data which could improve our understanding of AGWs propagation direction and source distribution in the MLT region. We have developed a new statistical analysis method for obtaining the power spectrum in the horizontal phase <span class="hlt">velocity</span> domain (phase <span class="hlt">velocity</span> spectrum), from airglow image data, so as to deal with huge amounts of imaging data obtained on different years and at various <span class="hlt">observation</span> sites, without bias caused by different event extraction criteria for the <span class="hlt">observer</span>. From a series of images projected onto the geographic coordinates, 3-D Fourier transform is applied and 3-D power spectrum in horizontal wavenumber and frequency domain is obtained. Then, it is converted into phase <span class="hlt">velocity</span> and frequency domain. Finally, the spectrum is integrated along the frequency for the range of interest and 2-D spectrum in horizontal phase <span class="hlt">velocity</span> is calculated. This method was applied to the data obtained at Syowa Station (69ºS, 40ºE), Antarctica, in 2011 and compared with a conventional event analysis in which the phase fronts were traced manually in order to estimate horizontal propagation characteristics. This comparison shows that our new method is adequate to deriving the horizontal phase <span class="hlt">velocity</span> characteristics of AGWs <span class="hlt">observed</span> by airglow imaging technique. Airglow imaging <span class="hlt">observation</span> has been operated with various sampling intervals. We also presents how the images with different sample interval should be treated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890064582&hterms=SPIRAL+MODEL&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DSPIRAL%2BMODEL','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890064582&hterms=SPIRAL+MODEL&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DSPIRAL%2BMODEL"><span>Cosmological <span class="hlt">velocity</span> correlations - <span class="hlt">Observations</span> and model predictions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gorski, Krzysztof M.; Davis, Marc; Strauss, Michael A.; White, Simon D. M.; Yahil, Amos</p> <p>1989-01-01</p> <p>By applying the present simple statistics for two-point cosmological peculiar <span class="hlt">velocity</span>-correlation measurements to the actual data sets of the Local Supercluster spiral galaxy of Aaronson et al. (1982) and the elliptical galaxy sample of Burstein et al. (1987), as well as to the <span class="hlt">velocity</span> field predicted by the distribution of IRAS galaxies, a coherence length of 1100-1600 km/sec is obtained. Coherence length is defined as that separation at which the correlations drop to half their zero-lag value. These results are compared with predictions from two models of large-scale structure formation: that of cold dark matter and that of baryon isocurvature proposed by Peebles (1980). N-body simulations of these models are performed to check the linear theory predictions and measure sampling fluctuations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997ApJ...481..267D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997ApJ...481..267D"><span>The Dynamics of M15: <span class="hlt">Observations</span> of the <span class="hlt">Velocity</span> Dispersion Profile and Fokker-Planck Models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dull, J. D.; Cohn, H. N.; Lugger, P. M.; Murphy, B. W.; Seitzer, P. O.; Callanan, P. J.; Rutten, R. G. M.; Charles, P. A.</p> <p>1997-05-01</p> <p>We report a new measurement of the <span class="hlt">velocity</span> dispersion profile within 1' (3 pc) of the center of the globular cluster M15 (NGC 7078), using long-slit spectra from the 4.2 m William Herschel Telescope at La Palma Observatory. We obtained spatially resolved spectra for a total of 23 slit positions during two <span class="hlt">observing</span> runs. During each run, a set of parallel slit positions was used to map out the central region of the cluster; the position angle used during the second run was orthogonal to that used for the first. The spectra are centered in wavelength near the Ca II infrared triplet at 8650 Å, with a spectral range of about 450 Å. We determined radial <span class="hlt">velocities</span> by cross-correlation techniques for 131 cluster members. A total of 32 stars were <span class="hlt">observed</span> more than once. Internal and external comparisons indicate a <span class="hlt">velocity</span> accuracy of about 4 km s-1. The <span class="hlt">velocity</span> dispersion profile rises from about σ = 7.2 +/- 1.4 km s-1 near 1' from the center of the cluster to σ = 13.9 +/- 1.8 km s-1 at 20". Inside of 20", the dispersion remains approximately constant at about 10.2 +/- 1.4 km s-1 with no evidence for a sharp rise near the center. This last result stands in contrast with that of Peterson, Seitzer, & Cudworth who found a central <span class="hlt">velocity</span> dispersion of 25 +/- 7 km s-1, based on a line-broadening measurement. Our <span class="hlt">velocity</span> dispersion profile is in good agreement with those determined in the recent studies of Gebhardt et al. and Dubath & Meylan. We have developed a new set of Fokker-Planck models and have fitted these to the surface brightness and <span class="hlt">velocity</span> dispersion profiles of M15. We also use the two measured millisecond pulsar accelerations as constraints. The best-fitting model has a mass function slope of x = 0.9 (where 1.35 is the slope of the Salpeter mass function) and a total mass of 4.9 × 105 M⊙. This model contains approximately 104 neutron stars (3% of the total mass), the majority of which lie within 6" (0.2 pc) of the cluster center. Since the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.6584T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.6584T"><span>Vertical rise <span class="hlt">velocity</span> of equatorial plasma bubbles estimated from Equatorial Atmosphere Radar (EAR) <span class="hlt">observations</span> and HIRB model simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tulasi Ram, S.; Ajith, K. K.; Yokoyama, T.; Yamamoto, M.; Niranjan, K.</p> <p>2017-06-01</p> <p>The vertical rise <span class="hlt">velocity</span> (Vr) and maximum altitude (Hm) of equatorial plasma bubbles (EPBs) were estimated using the two-dimensional fan sector maps of 47 MHz Equatorial Atmosphere Radar (EAR), Kototabang, during May 2010 to April 2013. A total of 86 EPBs were <span class="hlt">observed</span> out of which 68 were postsunset EPBs and remaining 18 EPBs were <span class="hlt">observed</span> around midnight hours. The vertical rise <span class="hlt">velocities</span> of the EPBs <span class="hlt">observed</span> around the midnight hours are significantly smaller ( 26-128 m/s) compared to those <span class="hlt">observed</span> in postsunset hours ( 45-265 m/s). Further, the vertical growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The three-dimensional numerical high-resolution bubble (HIRB) model with varying background conditions are employed to investigate the possible factors that control the vertical rise <span class="hlt">velocity</span> and maximum attainable altitudes of EPBs. The estimated rise <span class="hlt">velocities</span> from EAR <span class="hlt">observations</span> at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model. The smaller vertical rise <span class="hlt">velocities</span> (Vr) and lower maximum altitudes (Hm) of EPBs during midnight hours are discussed in terms of weak polarization electric fields within the bubble due to weaker background electric fields and reduced background ion density levels.<abstract type="synopsis"><title type="main">Plain Language SummaryEquatorial plasma bubbles are plasma density irregularities in the ionosphere. The radio waves passing through these irregular density structures undergo severe degradation/scintillation that could cause severe disruption of satellite-based communication and augmentation systems such as GPS navigation. These bubbles develop at geomagnetic equator, grow vertically, and elongate along the field lines to latitudes away from the equator. The knowledge on bubble rise <span class="hlt">velocities</span> and their</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H43B1631H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H43B1631H"><span>Study on of Seepage Flow <span class="hlt">Velocity</span> in Sand Layer Profile as Affected by Water Depth and Slope Gradience</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, Z.; Chen, X.</p> <p>2017-12-01</p> <p>BACKGROUND: The subsurface water flow <span class="hlt">velocity</span> is of great significance in understanding the hydrodynamic characteristics of soil seepage and the influence of interaction between seepage flow and surface runoff on the soil erosion and sediment transport process. OBJECTIVE: To propose a visualized method and equipment for determining the seepage flow <span class="hlt">velocity</span> and measuring the actual flow <span class="hlt">velocity</span> and Darcy <span class="hlt">velocity</span> as well as the relationship between them.METHOD: A transparent organic glass tank is used as the test soil tank, the white river sand is used as the seepage test material and the fluorescent dye is used as the indicator for tracing water flow, so as to determine the thickness and <span class="hlt">velocity</span> of water flow in a visualized way. Water is supplied at the same flow rate (0.84 L h-1) to the three parts with an interval of 1m at the bottom of the soil tank and the pore water <span class="hlt">velocity</span> and the thickness of each water layer are determined under four <span class="hlt">gradient</span> conditions. The Darcy <span class="hlt">velocity</span> of each layer is calculated according to the water supply flow and the discharge section area. The effective discharge flow pore is estimated according to the moisture content and porosity and then the relationship between Darcy <span class="hlt">velocity</span> and the measured <span class="hlt">velocity</span> is calculated based on the water supply flow and the water layer thickness, and finally the correctness of the calculation results is verified. RESULTS: According to the <span class="hlt">velocity</span> calculation results, Darcy <span class="hlt">velocity</span> increases significantly with the increase of <span class="hlt">gradient</span>; in the sand layer profile, the flow <span class="hlt">velocity</span> of pore water at different depths increases with the increase of <span class="hlt">gradient</span>; under the condition of the same <span class="hlt">gradient</span>, the lower sand layer has the maximum flow <span class="hlt">velocity</span> of pore water. The air-filled porosity of sand layer determines the proportional relationship between Darcy <span class="hlt">velocity</span> and pore flow <span class="hlt">velocity</span>. CONCLUSIONS: The actual flow <span class="hlt">velocity</span> and Darcy <span class="hlt">velocity</span> can be measured by a visualized method and the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1989/4090/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1989/4090/report.pdf"><span>Accuracy of acoustic <span class="hlt">velocity</span> metering systems for measurement of low <span class="hlt">velocity</span> in open channels</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Laenen, Antonius; Curtis, R. E.</p> <p>1989-01-01</p> <p>Acoustic <span class="hlt">velocity</span> meter (AVM) accuracy depends on equipment limitations, the accuracy of acoustic-path length and angle determination, and the stability of the mean <span class="hlt">velocity</span> to acoustic-path <span class="hlt">velocity</span> relation. Equipment limitations depend on path length and angle, transducer frequency, timing oscillator frequency, and signal-detection scheme. Typically, the <span class="hlt">velocity</span> error from this source is about +or-1 to +or-10 mms/sec. Error in acoustic-path angle or length will result in a proportional measurement bias. Typically, an angle error of one degree will result in a <span class="hlt">velocity</span> error of 2%, and a path-length error of one meter in 100 meter will result in an error of 1%. Ray bending (signal refraction) depends on path length and density <span class="hlt">gradients</span> present in the stream. Any deviation from a straight acoustic path between transducer will change the unique relation between path <span class="hlt">velocity</span> and mean <span class="hlt">velocity</span>. These deviations will then introduce error in the mean <span class="hlt">velocity</span> computation. Typically, for a 200-meter path length, the resultant error is less than one percent, but for a 1,000 meter path length, the error can be greater than 10%. Recent laboratory and field tests have substantiated assumptions of equipment limitations. Tow-tank tests of an AVM system with a 4.69-meter path length yielded an average standard deviation error of 9.3 mms/sec, and the field tests of an AVM system with a 20.5-meter path length yielded an average standard deviation error of a 4 mms/sec. (USGS)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/686875-velocity-jump-instabilities-hele-shaw-flow-associating-polymer-solutions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/686875-velocity-jump-instabilities-hele-shaw-flow-associating-polymer-solutions"><span><span class="hlt">Velocity</span>-jump instabilities in Hele-Shaw flow of associating polymer solutions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Vlad, D.H.; Ignes-Mullol, J.; Maher, J.V.</p> <p></p> <p>We study fracturelike flow instabilities that arise when water is injected into a Hele-Shaw cell filled with aqueous solutions of associating polymers. We explore various polymer architectures, molecular weights, and solution concentrations. Simultaneous measurements of the finger tip <span class="hlt">velocity</span> and of the pressure at the injection point allow us to describe the dynamics of the finger in terms of the {open_quotes}finger mobility,{close_quotes} which relates the <span class="hlt">velocity</span> to the pressure <span class="hlt">gradient</span>. The flow discontinuities, characterized by jumps in the finger tip <span class="hlt">velocity</span>, which are <span class="hlt">observed</span> in experiments with some of the polymer solutions, can be modeled by using a nonmonotonic dependencemore » between a characteristic shear stress and the shear rate at the tip of the finger. A simple model, which is based on a viscosity function containing both a Newtonian and a non-Newtonian component, and which predicts nonmonotonic regions when the non-Newtonian component of the viscosity dominates, is shown to agree with the experimental data. {copyright} {ital 1999} {ital The American Physical Society}« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994MsT.........17M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994MsT.........17M"><span>Compressible turbulence measurements in a supersonic boundary layer including favorable pressure <span class="hlt">gradient</span> effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miller, Raymond S.</p> <p>1994-12-01</p> <p>The effect of a favorable pressure <span class="hlt">gradient</span> on the turbulent flow structure in a Mach 2.9 boundary layer (Re/m approximately equal to 1.5 x 10(exp 7)) is investigated experimentally. Conventional flow and hot film measurements of turbulent fluctuation properties have been made upstream of and along an expansion ramp. Upstream measurements were taken in a zero pressure <span class="hlt">gradient</span> boundary layer 44 cm from the nozzle throat in a 6.35 cm square test section. Measurements are obtained in the boundary layer, above the expansion ramp, 71.5 cm from the nozzle throat. Mean flow and turbulent flow characteristics are measured in all three dimensions. Comparisons are made between data obtained using single and multiple-overheat cross-wire anemometry as well as conventional mean flow probes. Conventional flow measurements were taken using a Pitot probe and a 10 degree cone static probe. Flow visualization was conducted via imaging techniques (Schlieren and shadowgraph photographs). Results suggest that compressibility effects, as seen through the density fluctuations in the Reynolds shear stress, are roughly 10% relative to the mean <span class="hlt">velocity</span> and are large relative to the <span class="hlt">velocity</span> fluctuations. This is also <span class="hlt">observed</span> in the total Reynolds shear stress; compressibility accounts for 50-75% of the total shear. This is particularly true in the favorable pressure <span class="hlt">gradient</span> region, where though the peak fluctuation intensities are diminished, the streamwise component of the mean flow is larger, hence the contribution of the compressibility term is significant in the Reynolds shear.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20030003640','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20030003640"><span>Shadowgraph Study of <span class="hlt">Gradient</span> Driven Fluctuations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cannell, David; Nikolaenko, Gennady; Giglio, Marzio; Vailati, Alberto; Croccolo, Fabrizio; Meyer, William</p> <p>2002-01-01</p> <p>A fluid or fluid mixture, subjected to a vertical temperature and/or concentration <span class="hlt">gradient</span> in a gravitational field, exhibits greatly enhanced light scattering at small angles. This effect is caused by coupling between the vertical <span class="hlt">velocity</span> fluctuations due to thermal energy and the vertically varying refractive index. Physically, small upward or downward moving regions will be displaced into fluid having a refractive index different from that of the moving region, thus giving rise to the enhanced scattering. The scattered intensity is predicted to vary with scattering wave vector q, as q(sup -4), for sufficiently large q, but the divergence is quenched by gravity at small q. In the absence of gravity, the long wavelength fluctuations responsible for the enhanced scattering are predicted to grow until limited by the sample dimensions. It is thus of interest to measure the mean-squared amplitude of such fluctuations in the microgravity environment for comparison with existing theory and ground based measurements. The relevant wave vectors are extremely small, making traditional low-angle light scattering difficult or impossible because of stray elastically scattered light generated by optical surfaces. An alternative technique is offered by the shadowgraph method, which is normally used to visualize fluid flows, but which can also serve as a quantitative tool to measure fluctuations. A somewhat novel shadowgraph apparatus and the necessary data analysis methods will be described. The apparatus uses a spatially coherent, but temporally incoherent, light source consisting of a super-luminescent diode coupled to a single-mode optical fiber in order to achieve extremely high spatial resolution, while avoiding effects caused by interference of light reflected from the various optical surfaces that are present when using laser sources. Results obtained for a critical mixture of aniline and cyclohexane subjected to a vertical temperature <span class="hlt">gradient</span> will be presented. The</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23432054','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23432054"><span>Analysis of the electrolyte convection inside the concentration boundary layer during structured electrodeposition of copper in high magnetic <span class="hlt">gradient</span> fields.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>König, Jörg; Tschulik, Kristina; Büttner, Lars; Uhlemann, Margitta; Czarske, Jürgen</p> <p>2013-03-19</p> <p>To experimentally reveal the correlation between electrodeposited structure and electrolyte convection induced inside the concentration boundary layer, a highly inhomogeneous magnetic field, generated by a magnetized Fe-wire, has been applied to an electrochemical system. The influence of Lorentz and magnetic field <span class="hlt">gradient</span> force to the local transport phenomena of copper ions has been studied using a novel two-component laser Doppler <span class="hlt">velocity</span> profile sensor. With this sensor, the electrolyte convection within 500 μm of a horizontally aligned cathode is presented. The electrode-normal two-component <span class="hlt">velocity</span> profiles below the electrodeposited structure show that electrolyte convection is induced and directed toward the rim of the Fe-wire. The measured deposited structure directly correlates to the <span class="hlt">observed</span> boundary layer flow. As the local concentration of Cu(2+) ions is enhanced due to the induced convection, maximum deposit thicknesses can be found at the rim of the Fe-wire. Furthermore, a complex boundary layer flow structure was determined, indicating that electrolyte convection of second order is induced. Moreover, the Lorentz force-driven convection rapidly vanishes, while the electrolyte convection induced by the magnetic field <span class="hlt">gradient</span> force is preserved much longer. The progress for research is the first direct experimental proof of the electrolyte convection inside the concentration boundary layer that correlates to the deposited structure and reveals that the magnetic field <span class="hlt">gradient</span> force is responsible for the <span class="hlt">observed</span> structuring effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C41B1190U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C41B1190U"><span>Seasonal and inter-annual variability in <span class="hlt">velocity</span> and frontal position of Siachen Glacier (Eastern Karakorum) using multi-satellite data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Usman, M.; Furuya, M.; Sakakibara, D.; Abe, T.</p> <p>2017-12-01</p> <p>The anomalous behavior of Karakorum glaciers is a hot topic of discussion in the scientific community. Siachen Glacier is one of the longest glaciers ( 75km) in Karakorum Range. This glacier is supposed to be a surge type but so far no studies have confirmed this claim. Detailed <span class="hlt">velocity</span> mapping of this glacier can possibly provide some clues about intra/inter-annual changes in <span class="hlt">velocity</span> and <span class="hlt">observed</span> terminus. Using L-band SAR data of ALOS-1/2, we applied the feature tracking technique (search patch of 128x128 pixels (range x azimuth) , sampling interval of 12x36 pixels) to derive <span class="hlt">velocity</span> changes; we used GAMMA software. The <span class="hlt">velocity</span> was calculated by following the parallel flow assumption. To calculate the local topographic <span class="hlt">gradient</span> unit vector, we used ASTER-GDEM. We also used optical images acquired by Landsat 5 Thematic Mapper (TM), the Landsat 7 Enhanced Thematic Mapper Plus (ETM+) to derive surface <span class="hlt">velocity</span>. The algorithm we used is Cross-Correlation in Frequency domain on Orientation images (CCF-O). The <span class="hlt">velocity</span> was finally calculated by setting a flow line and averaging over the area of 200x200m2. The results indicate seasonal speed up signals that modulate inter-annually from 1999 to 2011, with slight or no change in the <span class="hlt">observed</span> frontal position. However, in ALOS-2 data, the `<span class="hlt">observed</span> terminus' seems to have been advancing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S21A0694S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S21A0694S"><span>Applications of seismic spatial wavefield <span class="hlt">gradient</span> and rotation data in exploration seismology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmelzbach, C.; Van Renterghem, C.; Sollberger, D.; Häusler, M.; Robertsson, J. O. A.</p> <p>2017-12-01</p> <p>Seismic spatial wavefield <span class="hlt">gradient</span> and rotation data have the potential to open up new ways to address long-standing problems in land-seismic exploration such as identifying and separating P-, S-, and surface waves. <span class="hlt">Gradient</span>-based acquisition and processing techniques could enable replacing large arrays of densely spaced receivers by sparse spatially-compact receiver layouts or even one single multicomponent station with dedicated instruments (e.g., rotational seismometers). Such approaches to maximize the information content of single-station recordings are also of significant interest for seismic measurements at sites with limited access such as boreholes, the sea bottom, and extraterrestrial seismology. Arrays of conventional three-component (3C) geophones enable measuring not only the particle <span class="hlt">velocity</span> in three dimensions but also estimating their spatial <span class="hlt">gradients</span>. Because the free-surface condition allows to express vertical derivatives in terms of horizontal derivatives, the full <span class="hlt">gradient</span> tensor and, hence, curl and divergence of the wavefield can be computed. In total, three particle <span class="hlt">velocity</span> components, three rotational components, and divergence, result seven-component (7C) seismic data. Combined particle <span class="hlt">velocity</span> and <span class="hlt">gradient</span> data can be used to isolate the incident P- or S-waves at the land surface or the sea bottom using filtering techniques based on the elastodynamic representation theorem. Alternatively, as only S-waves exhibit rotational motion, rotational measurements can directly be used to identify S-waves. We discuss the derivations of the <span class="hlt">gradient</span>-based filters as well as their application to synthetic and field data, demonstrating that rotational data can be of particular interest to S-wave reflection and P-to-S-wave conversion imaging. The concept of array-derived <span class="hlt">gradient</span> estimation can be extended to source arrays as well. Therefore, source arrays allow us to emulate rotational (curl) and dilatational (divergence) sources. Combined with 7C</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C51D..05M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51D..05M"><span>Quantifying seasonal <span class="hlt">velocity</span> at Khumbu Glacier, Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miles, E.; Quincey, D. J.; Miles, K.; Hubbard, B. P.; Rowan, A. V.</p> <p>2017-12-01</p> <p>While the low-<span class="hlt">gradient</span> debris-covered tongues of many Himalayan glaciers exhibit low surface <span class="hlt">velocities</span>, quantifying ice flow and its variation through time remains a key challenge for studies aimed at determining the long-term evolution of these glaciers. Recent work has suggested that glaciers in the Everest region of Nepal may show seasonal variability in surface <span class="hlt">velocity</span>, with ice flow peaking during the summer as monsoon precipitation provides hydrological inputs and thus drives changes in subglacial drainage efficiency. However, satellite and aerial <span class="hlt">observations</span> of glacier <span class="hlt">velocity</span> during the monsoon are greatly limited due to cloud cover. Those that do exist do not span the period over which the most dynamic changes occur, and consequently short-term (i.e. daily) changes in flow, as well as the evolution of ice dynamics through the monsoon period, remain poorly understood. In this study, we combine field and remote (satellite image) <span class="hlt">observations</span> to create a multi-temporal, 3D synthesis of ice deformation rates at Khumbu Glacier, Nepal, focused on the 2017 monsoon period. We first determine net annual and seasonal surface displacements for the whole glacier based on Landsat-8 (OLI) panchromatic data (15m) processed with ImGRAFT. We integrate inclinometer <span class="hlt">observations</span> from three boreholes drilled by the EverDrill project to determine cumulative deformation at depth, providing a 3D perspective and enabling us to assess the role of basal sliding at each site. We additionally analyze high-frequency on-glacier L1 GNSS data from three sites to characterize variability within surface deformation at sub-seasonal timescales. Finally, each dataset is validated against repeat-dGPS <span class="hlt">observations</span> at gridded points in the vicinity of the boreholes and GNSS dataloggers. These datasets complement one another to infer thermal regime across the debris-covered ablation area of the glacier, and emphasize the seasonal and spatial variability of ice deformation for glaciers in High</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025436','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025436"><span>The relationship between the instantaneous <span class="hlt">velocity</span> field and the rate of moment release in the lithosphere</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pollitz, F.F.</p> <p>2003-01-01</p> <p>Instantaneous <span class="hlt">velocity</span> <span class="hlt">gradients</span> within the continental lithosphere are often related to the tectonic driving forces. This relationship is direct if the forces are secular, as for the case of loading of a locked section of a subduction interface by the downgoing plate. If the forces are static, as for the case of lateral variations in gravitational potential energy, then <span class="hlt">velocity</span> <span class="hlt">gradients</span> can be produced only if the lithosphere has, on average, zero strength. The static force model may be related to the long-term <span class="hlt">velocity</span> field but not the instantaneous <span class="hlt">velocity</span> field (typically measured geodetically over a period of several years) because over short time intervals the upper lithosphere behaves elastically. In order to describe both the short- and long-term behaviour of an (elastic) lithosphere-(viscoelastic) asthenosphere system in a self-consistent manner, I construct a deformation model termed the expected interseismic <span class="hlt">velocity</span> (EIV) model. Assuming that the lithosphere is populated with faults that rupture continually, each with a definite mean recurrence time, and that the Earth is well approximated as a linear elastic-viscoelastic coupled system, I derive a simple relationship between the instantaneous <span class="hlt">velocity</span> field and the average rate of moment release in the lithosphere. Examples with synthetic fault networks demonstrate that <span class="hlt">velocity</span> <span class="hlt">gradients</span> in actively deforming regions may to a large extent be the product of compounded viscoelastic relaxation from past earthquakes on hundreds of faults distributed over large ( ≥106 km2) areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/twri/twri3-a17/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/twri/twri3-a17/"><span>Acoustic <span class="hlt">velocity</span> meter systems</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Laenen, Antonius</p> <p>1985-01-01</p> <p>Acoustic <span class="hlt">velocity</span> meter (AVM) systems operate on the principles that the point-to-point upstream traveltime of an acoustic pulse is longer than the downstream traveltime and that this difference in traveltime can be accurately measured by electronic devices. An AVM system is capable of recording water <span class="hlt">velocity</span> (and discharge) under a wide range of conditions, but some constraints apply: 1. Accuracy is reduced and performance is degraded if the acoustic path is not a continuous straight line. The path can be bent by reflection if it is too close to a stream boundary or by refraction if it passes through density <span class="hlt">gradients</span> resulting from variations in either water temperature or salinity. For paths of less than 100 m, a temperature <span class="hlt">gradient</span> of 0.1' per meter causes signal bending less than 0.6 meter at midchannel, and satisfactory <span class="hlt">velocity</span> results can be obtained. Reflection from stream boundaries can cause signal cancellation if boundaries are too close to signal path. 2. Signal strength is attenuated by particles or bubbles that absorb, spread, or scatter sound. The concentration of particles or bubbles that can be tolerated is a function of the path length and frequency of the acoustic signal. 3. Changes in streamline orientation can affect system accuracy if the variability is random. 4. Errors relating to signal resolution are much larger for a single threshold detection scheme than for multiple threshold schemes. This report provides methods for computing the effect of various conditions on the accuracy of a record obtained from an AVM. The equipment must be adapted to the site. Field reconnaissance and preinstallation analysis to detect possible problems are critical for proper installation and operation of an AVM system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19920015713','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19920015713"><span><span class="hlt">Observations</span> on computational methodologies for use in large-scale, <span class="hlt">gradient</span>-based, multidisciplinary design incorporating advanced CFD codes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Newman, P. A.; Hou, G. J.-W.; Jones, H. E.; Taylor, A. C., III; Korivi, V. M.</p> <p>1992-01-01</p> <p>How a combination of various computational methodologies could reduce the enormous computational costs envisioned in using advanced CFD codes in <span class="hlt">gradient</span> based optimized multidisciplinary design (MdD) procedures is briefly outlined. Implications of these MdD requirements upon advanced CFD codes are somewhat different than those imposed by a single discipline design. A means for satisfying these MdD requirements for <span class="hlt">gradient</span> information is presented which appear to permit: (1) some leeway in the CFD solution algorithms which can be used; (2) an extension to 3-D problems; and (3) straightforward use of other computational methodologies. Many of these <span class="hlt">observations</span> have previously been discussed as possibilities for doing parts of the problem more efficiently; the contribution here is <span class="hlt">observing</span> how they fit together in a mutually beneficial way.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028027','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028027"><span>UHF RiverSonde <span class="hlt">observations</span> of water surface <span class="hlt">velocity</span> at Threemile Slough, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Teague, C.C.; Barrick, D.E.; Lilleboe, P.M.; Cheng, R.T.; Ruhl, C.A.</p> <p>2005-01-01</p> <p>A UHF RiverSonde system, operating near 350 MHz, has been in operation at Threemile Slough in central California, USA since September 2004. The water in the slough is dominated by tidal effects, with flow reversals four times a day and a peak <span class="hlt">velocity</span> of about 0.8 m/s in each direction. Water level and water <span class="hlt">velocity</span> are continually measured by the U. S. Geological Survey at the experiment site. The <span class="hlt">velocity</span> is measured every 15 minutes by an ultrasonic <span class="hlt">velocity</span> meter (UVM) which determines the water <span class="hlt">velocity</span> from two-way acoustic propagation time-difference measurements made across the channel. The RiverSonde also measures surface <span class="hlt">velocity</span> every 15 minutes using radar resonant backscatter techniques. <span class="hlt">Velocity</span> and water level data are retrieved through a radio data link and a wideband internet connection. Over a period of several months, the radar-derived mean surface <span class="hlt">velocity</span> has been very highly correlated with the UVM index <span class="hlt">velocity</span> several meters below the surface, with a coefficient of determination R2 of 0.976 and an RMS difference of less than 10 cm/s. The wind has a small but measurable effect on the <span class="hlt">velocities</span> measured by both instruments. In addition to the mean surface <span class="hlt">velocity</span> across the channel, the RiverSonde system provides an estimate of the cross-channel variation of the surface <span class="hlt">velocity</span>. ?? 2005 IEEE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23366530','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23366530"><span>ECG denoising using angular <span class="hlt">velocity</span> as a state and an <span class="hlt">observation</span> in an Extended Kalman Filter framework.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Akhbari, Mahsa; Shamsollahi, Mohammad B; Jutten, Christian; Coppa, Bertrand</p> <p>2012-01-01</p> <p>In this paper an efficient filtering procedure based on Extended Kalman Filter (EKF) has been proposed. The method is based on a modified nonlinear dynamic model, previously introduced for the generation of synthetic ECG signals. The proposed method considers the angular <span class="hlt">velocity</span> of ECG signal, as one of the states of an EKF. We have considered two cases for <span class="hlt">observation</span> equations, in one case we have assumed a corresponding <span class="hlt">observation</span> to angular <span class="hlt">velocity</span> state and in the other case, we have not assumed any <span class="hlt">observations</span> for it. Quantitative evaluation of the proposed algorithm on the MIT-BIH Normal Sinus Rhythm Database (NSRDB) shows that an average SNR improvement of 8 dB is achieved for an input signal of -4 dB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890045672&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D90%26Ntt%3Dlazarus','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890045672&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D90%26Ntt%3Dlazarus"><span>Pioneer and Voyager <span class="hlt">observations</span> of the solar wind at large heliocentric distances and latitudes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gazis, P. R.; Mihalov, J. D.; Barnes, A.; Lazarus, A. J.; Smith, E. J.</p> <p>1989-01-01</p> <p>Data obtained from the electrostatic analyzers aboard the Pioneer 10 and 11 spacecraft and from the Faraday cup aboard Voyager 2 were used to study spatial <span class="hlt">gradients</span> in the distant solar wind. Prior to mid-1985, both spacecraft <span class="hlt">observed</span> nearly identical solar wind structures. After day 150 of 1985, the <span class="hlt">velocity</span> structure at Voyager 2 became flatter, and the Voyager 2 <span class="hlt">velocities</span> were smaller than those <span class="hlt">observed</span> by Pioneer 11. It is suggested that these changes in the solar wind at low latitudes may be related to a change which occurred in the coronal hole structure in early 1985.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.2467F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.2467F"><span>Global view of the E region irregularity and convection <span class="hlt">velocities</span> in the high-latitude Southern Hemisphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Forsythe, Victoriya V.; Makarevich, Roman A.</p> <p>2017-02-01</p> <p>Occurrence of the E region plasma irregularities is investigated using two Super Dual Auroral Radar Network (SuperDARN) South Pole (SPS) and Zhongshan (ZHO) radars that sample the same magnetic latitude deep within the high-latitude plasma convection pattern but from two opposite directions. It is shown that the SPS and ZHO <span class="hlt">velocity</span> distributions and their variations with the magnetic local time are different, with each distribution being asymmetric; i.e., a particular <span class="hlt">velocity</span> polarity is predominant. This asymmetry in the E region <span class="hlt">velocity</span> distribution is associated with the bump-on-tail of the distribution near the nominal ion acoustic speed Cs that is most likely due to the Farley-Buneman instability (FBI) echoes or an inflection point of the distribution below nominal Cs that is most likely due to the <span class="hlt">gradient</span> drift instability echoes. In contrast, the distribution of the convection <span class="hlt">velocity</span> component was found to be symmetric, i.e., with no bump-on-tail or an inflection point, but with a bias (i.e., uniform shift) toward a particular polarity. It is demonstrated that the asymmetry in the convection pattern between the eastward and westward zonal components is unexpectedly strong, with the westward zonal component being predominant, especially at lower latitudes, while also exhibiting a strong interplanetary magnetic field By dependence. The <span class="hlt">observations</span> are consistent with the notion that the asymmetry in the E region <span class="hlt">velocity</span> distribution is highly sensitive to the bias in the convection component caused by the zonal convection component asymmetry and that the bump-on-tail or inflection point features may also depend on the irregularity height and the presence of strong density <span class="hlt">gradients</span> modifying the FBI threshold value.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29846183','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29846183"><span>Thermal <span class="hlt">gradients</span> for the stabilization of a single domain wall in magnetic nanowires.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mejía-López, J; Velásquez, E A; Mazo-Zuluaga, J; Altbir, D</p> <p>2018-08-24</p> <p>By means of Monte Carlo simulations we studied field driven nucleation and propagation of transverse domain walls (DWs) in magnetic nanowires subjected to temperature <span class="hlt">gradients</span>. Simulations identified the existence of critical thermal <span class="hlt">gradients</span> that allow the existence of reversal processes driven by a single DW. Critical thermal <span class="hlt">gradients</span> depend on external parameters such as temperature, magnetic field and wire length, and can be experimentally obtained through the measurement of the mean <span class="hlt">velocity</span> of the magnetization reversal as a function of the temperature <span class="hlt">gradient</span>. Our results show that temperature <span class="hlt">gradients</span> provide a high degree of control over DW propagation, which is of great importance for technological applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70195193','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70195193"><span>The suitability of using dissolved gases to determine groundwater discharge to high <span class="hlt">gradient</span> streams</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gleeson, Tom; Manning, Andrew H.; Popp, Andrea; Zane, Mathew; Clark, Jordan F.</p> <p>2018-01-01</p> <p>Determining groundwater discharge to streams using dissolved gases is known to be useful over a wide range of streamflow rates but the suitability of dissolved gas methods to determine discharge rates in high <span class="hlt">gradient</span> mountain streams has not been sufficiently tested, even though headwater streams are critical as ecological habitats and water resources. The aim of this study is to test the suitability of using dissolved gases to determine groundwater discharge rates to high <span class="hlt">gradient</span> streams by field experiments in a well-characterized, high <span class="hlt">gradient</span> mountain stream and a literature review. At a reach scale (550 m) we combined stream and groundwater radon activity measurements with an in-stream SF6 tracer test. By means of numerical modeling we determined gas exchange <span class="hlt">velocities</span> and derived very low groundwater discharge rates (∼15% of streamflow). These groundwater discharge rates are below the uncertainty range of physical streamflow measurements and consistent with temperature, specific conductance and streamflow measured at multiple locations along the reach. At a watershed-scale (4 km), we measured CFC-12 and δ18O concentrations and determined gas exchange <span class="hlt">velocities</span> and groundwater discharge rates with the same numerical model. The groundwater discharge rates along the 4 km stream reach were highly variable, but were consistent with the values derived in the detailed study reach. Additionally, we synthesized literature values of gas exchange <span class="hlt">velocities</span> for different stream <span class="hlt">gradients</span> which show an empirical relationship that will be valuable in planning future dissolved gas studies on streams with various <span class="hlt">gradients</span>. In sum, we show that multiple dissolved gas tracers can be used to determine groundwater discharge to high <span class="hlt">gradient</span> mountain streams from reach to watershed scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..557..561G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..557..561G"><span>The suitability of using dissolved gases to determine groundwater discharge to high <span class="hlt">gradient</span> streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gleeson, Tom; Manning, Andrew H.; Popp, Andrea; Zane, Matthew; Clark, Jordan F.</p> <p>2018-02-01</p> <p>Determining groundwater discharge to streams using dissolved gases is known to be useful over a wide range of streamflow rates but the suitability of dissolved gas methods to determine discharge rates in high <span class="hlt">gradient</span> mountain streams has not been sufficiently tested, even though headwater streams are critical as ecological habitats and water resources. The aim of this study is to test the suitability of using dissolved gases to determine groundwater discharge rates to high <span class="hlt">gradient</span> streams by field experiments in a well-characterized, high <span class="hlt">gradient</span> mountain stream and a literature review. At a reach scale (550 m) we combined stream and groundwater radon activity measurements with an in-stream SF6 tracer test. By means of numerical modeling we determined gas exchange <span class="hlt">velocities</span> and derived very low groundwater discharge rates (∼15% of streamflow). These groundwater discharge rates are below the uncertainty range of physical streamflow measurements and consistent with temperature, specific conductance and streamflow measured at multiple locations along the reach. At a watershed-scale (4 km), we measured CFC-12 and δ18O concentrations and determined gas exchange <span class="hlt">velocities</span> and groundwater discharge rates with the same numerical model. The groundwater discharge rates along the 4 km stream reach were highly variable, but were consistent with the values derived in the detailed study reach. Additionally, we synthesized literature values of gas exchange <span class="hlt">velocities</span> for different stream <span class="hlt">gradients</span> which show an empirical relationship that will be valuable in planning future dissolved gas studies on streams with various <span class="hlt">gradients</span>. In sum, we show that multiple dissolved gas tracers can be used to determine groundwater discharge to high <span class="hlt">gradient</span> mountain streams from reach to watershed scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1984GeoRL..11..761H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1984GeoRL..11..761H"><span>Zonal pressure <span class="hlt">gradient</span>, <span class="hlt">velocity</span> and transport in the Atlantic Equatorial Undercurrent from focal cruises (July 1982-February 1984)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hisard, Philippe; Hénin, Christian</p> <p></p> <p>The zonal pressure <span class="hlt">gradient</span> (ZPG) along the Atlantic equator and the Equatorial Undercurrent (EUC) transport are discussed for four cruises representative of each season. A very clear sea surface slope reversal occurred in the eastern area during autumn as far west as 14°W. An early onset of the equatorial thermocline rising was evident during spring 1983. An eastward equatorial surface jet clearly distinct from the EUC was <span class="hlt">observed</span> at 35°W and 29°W. The greatest ZPG but the lowest EUC transport were <span class="hlt">observed</span> during summer 1983. A nearly total absence of the ZPG and a large surfacing of the EUC as far as 10°W characterized the 1984 winter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JGR....99.8873G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JGR....99.8873G"><span>Coupling of microprocesses and macroprocesses due to <span class="hlt">velocity</span> shear: An application to the low-altitude ionosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ganguli, G.; Keskinen, M. J.; Romero, H.; Heelis, R.; Moore, T.; Pollock, C.</p> <p>1994-05-01</p> <p>Recent <span class="hlt">observations</span> indicate that low-altitude (below 1500 km) ion energization and thermal ion upwelling are colocated in the convective flow reversal region. In this region the convective <span class="hlt">velocity</span> V(sub perpendicular) is generally small but spatial <span class="hlt">gradients</span> in V(sub perpendicular) can be large. As a result, Joule heating is small. The <span class="hlt">observed</span> high level of ion heating (few electron volts or more) cannot be explained by classical Joule heating alone but requires additional heating sources such as plasma waves. At these lower altitudes, sources of free energy are not obvious and hence the nature of ion energization remains ill understood. The high degree of correlation of ion heating with shear in the convective <span class="hlt">velocity</span> (Tsunoda et al., 1989) is suggestive of an important role of <span class="hlt">velocity</span> shear in this phenomenon. We provide more recent evidence for this correlation and show that even a small amount of <span class="hlt">velocity</span> shear in the transverse flow is sufficient to excite a large-scale Kelvin-Helmholtz mode, which can nonlinearly steepen and give rise to highly stressed regions of strongly sheared flows. Futhermore, these stressed regions of strongly sheared flows may seed plasma waves in the range of ion cyclotron to lower hybrid frequencies, which are potential sources for ion heating. This novle two-step mechanism for ion energization is applied to typical <span class="hlt">observations</span> of low-altitude thermal ion upwelling events.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EPJP..131..332S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EPJP..131..332S"><span>Thermally developing MHD peristaltic transport of nanofluids with <span class="hlt">velocity</span> and thermal slip effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sher Akbar, Noreen; Bintul Huda, A.; Tripathi, D.</p> <p>2016-09-01</p> <p>We investigate the <span class="hlt">velocity</span> slip and thermal slip effects on peristaltically driven thermal transport of nanofluids through the vertical parallel plates under the influence of transverse magnetic field. The wall surface is propagating with sinusoidal wave <span class="hlt">velocity</span> c. The flow characteristics are governed by the mass, momentum and energy conservation principle. Low Reynolds number and large wavelength approximations are taken into consideration to simplify the non-linear terms. Analytical solutions for axial <span class="hlt">velocity</span>, temperature field, pressure <span class="hlt">gradient</span> and stream function are obtained under certain physical boundary conditions. Two types of nanoparticles, SiO2 and Ag, are considered for analysis with water as base fluid. This is the first article in the literature that discusses the SiO2 and Ag nanoparticles for a peristaltic flow with variable viscosity. The effects of physical parameters on <span class="hlt">velocity</span>, temperature, pressure and trapping are discussed. A comparative study of SiO2 nanofluid, Ag nanofluid and pure water is also presented. This model is applicable in biomedical engineering to make thermal peristaltic pumps and other pumping devices like syringe pumps, etc. It is <span class="hlt">observed</span> that pressure for pure water is maximum and pressure for Ag nanofluid is minimum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22370429-measurements-outflow-velocities-disk-plumes-from-eis-hinode-observations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22370429-measurements-outflow-velocities-disk-plumes-from-eis-hinode-observations"><span>Measurements of outflow <span class="hlt">velocities</span> in on-disk plumes from EIS/Hinode <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Fu, Hui; Xia, Lidong; Li, Bo</p> <p>2014-10-20</p> <p>The contribution of plumes to the solar wind has been subject to hot debate in the past decades. The EUV Imaging Spectrometer (EIS) on board Hinode provides a unique means to deduce outflow <span class="hlt">velocities</span> at coronal heights via direct Doppler shift measurements of coronal emission lines. Such direct Doppler shift measurements were not possible with previous spectrometers. We measure the outflow <span class="hlt">velocity</span> at coronal heights in several on-disk long-duration plumes, which are located in coronal holes (CHs) and show significant blueshifts throughout the entire <span class="hlt">observational</span> period. In one case, a plume is measured four hours apart. The deduced outflow velocitiesmore » are consistent, suggesting that the flows are quasi-steady. Furthermore, we provide an outflow <span class="hlt">velocity</span> profile along the plumes, finding that the <span class="hlt">velocity</span> corrected for the line-of-sight effect can reach 10 km s{sup –1} at 1.02 R {sub ☉}, 15 km s{sup –1} at 1.03 R {sub ☉}, and 25 km s{sup –1} at 1.05 R {sub ☉}. This clear signature of steady acceleration, combined with the fact that there is no significant blueshift at the base of plumes, provides an important constraint on plume models. At the height of 1.03 R {sub ☉}, EIS also deduced a density of 1.3 × 10{sup 8} cm{sup –3}, resulting in a proton flux of about 4.2 × 10{sup 9} cm{sup –2} s{sup –1} scaled to 1 AU, which is an order of magnitude higher than the proton input to a typical solar wind if a radial expansion is assumed. This suggests that CH plumes may be an important source of the solar wind.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1343410-main-ion-intrinsic-toroidal-rotation-profile-driven-residual-stress-torque-from-ion-temperature-gradient-turbulence-diii-tokamak','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1343410-main-ion-intrinsic-toroidal-rotation-profile-driven-residual-stress-torque-from-ion-temperature-gradient-turbulence-diii-tokamak"><span>Main-ion intrinsic toroidal rotation profile driven by residual stress torque from ion temperature <span class="hlt">gradient</span> turbulence in the DIII-D tokamak</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Grierson, B. A.; Wang, W. X.; Ethier, S.; ...</p> <p>2017-01-06</p> <p>Intrinsic toroidal rotation of the deuterium main ions in the core of the DIII-D tokamak is <span class="hlt">observed</span> to transition from flat to hollow, forming an off-axis peak, above a threshold level of direct electron heating. Nonlinear gyrokinetic simulations show that the residual stress associated with electrostatic ion temperature <span class="hlt">gradient</span> turbulence possesses the correct radial location and stress structure to cause the <span class="hlt">observed</span> hollow rotation profile. Residual stress momentum flux in the gyrokinetic simulations is balanced by turbulent momentum diffusion, with negligible contributions from turbulent pinch. Finally, the prediction of the <span class="hlt">velocity</span> profile by integrating the momentum balance equation produces amore » rotation profile that qualitatively and quantitatively agrees with the measured main-ion profile, demonstrating that fluctuation-induced residual stress can drive the <span class="hlt">observed</span> intrinsic <span class="hlt">velocity</span> profile.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvF...3a2601W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvF...3a2601W"><span>Derivation of Zagarola-Smits scaling in zero-pressure-<span class="hlt">gradient</span> turbulent boundary layers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wei, Tie; Maciel, Yvan</p> <p>2018-01-01</p> <p>This Rapid Communication derives the Zagarola-Smits scaling directly from the governing equations for zero-pressure-<span class="hlt">gradient</span> turbulent boundary layers (ZPG TBLs). It has long been <span class="hlt">observed</span> that the scaling of the mean streamwise <span class="hlt">velocity</span> in turbulent boundary layer flows differs in the near surface region and in the outer layer. In the inner region of small-<span class="hlt">velocity</span>-defect boundary layers, it is generally accepted that the proper <span class="hlt">velocity</span> scale is the friction <span class="hlt">velocity</span>, uτ, and the proper length scale is the viscous length scale, ν /uτ . In the outer region, the most generally used length scale is the boundary layer thickness, δ . However, there is no consensus on <span class="hlt">velocity</span> scales in the outer layer. Zagarola and Smits [ASME Paper No. FEDSM98-4950 (1998)] proposed a <span class="hlt">velocity</span> scale, U ZS=(δ1/δ ) U∞ , where δ1 is the displacement thickness and U∞ is the freestream <span class="hlt">velocity</span>. However, there are some concerns about Zagarola-Smits scaling due to the lack of a theoretical base. In this paper, the Zagarola-Smits scaling is derived directly from a combination of integral, similarity, and order-of-magnitude analysis of the mean continuity equation. The analysis also reveals that V∞, the mean wall-normal <span class="hlt">velocity</span> at the edge of the boundary layer, is a proper scale for the mean wall-normal <span class="hlt">velocity</span> V . Extending the analysis to the streamwise mean momentum equation, we find that the Reynolds shear stress in ZPG TBLs scales as U∞V∞ in the outer region. This paper also provides a detailed analysis of the mass and mean momentum balance in the outer region of ZPG TBLs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20050180394','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20050180394"><span>Retrieval of Raindrop Size Distribution, Vertical Air <span class="hlt">Velocity</span> and Water Vapor Attenuation Using Dual-Wavelength Doppler Radar <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heymsfield, Gerald M.; Tian, Lin; Li, Lihua; Srivastava, C.</p> <p>2005-01-01</p> <p>Two techniques for retrieving the slope and intercept parameters of an assumed exponential raindrop size distribution (RSD), vertical air <span class="hlt">velocity</span>, and attenuation by precipitation and water vapor in light stratiform rain using <span class="hlt">observations</span> by airborne, nadir looking dual-wavelength (X-band, 3.2 cm and W-band, 3.2 mm) radars are presented. In both techniques, the slope parameter of the RSD and the vertical air <span class="hlt">velocity</span> are retrieved using only the mean Doppler <span class="hlt">velocities</span> at the two wavelengths. In the first method, the intercept of the RSD is estimated from the <span class="hlt">observed</span> reflectivity at the longer wavelength assuming no attenuation at that wavelength. The attenuation of the shorter wavelength radiation by precipitation and water vapor are retrieved using the <span class="hlt">observed</span> reflectivity at the shorter wavelength. In the second technique, it is assumed that the longer wavelength suffers attenuation only in the melting band. Then, assuming a distribution of water vapor, the melting band attenuation at both wavelengths and the rain attenuation at the shorter wavelength are retrieved. Results of the retrievals are discussed and several physically meaningful results are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMOS21A1660N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMOS21A1660N"><span>Radiosonde <span class="hlt">observational</span> evidence of the influence of extreme SST <span class="hlt">gradient</span> upon atmospheric meso-scale circulation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nishikawa, H.; Tachibana, Y.; Udagawa, Y.</p> <p>2012-12-01</p> <p>Although the influence of the anomalous midlatitude SST upon atmospheric local circulation has been getting common in particular over the Kuroshio and the Gulf Stream regions, <span class="hlt">observational</span> studies on the influence of the Okhotsk Sea, which is to the north of the Kuroshio, upon the atmospheric local circulation is much less than those of the Kuroshio. The climate of the Okhotsk SST is very peculiar. Extremely cold SST spots, whose summertime SST is lower than 5 Celsius degrees, are formed around Kuril Islands. Because SSTs are generally determined by local air-sea interaction as well as temperature advection, it is very difficult to isolate only the oceanic influence upon the atmosphere. The SST in this cold spot is, however, dominated by the tidal mixing, which is independent of the atmospheric processes. This unique condition may ease the account for the oceanic influence only. Although the SST environment of the Okhotsk Sea is good for understanding the oceanic influence upon the atmosphere, only a few studies has been executed in this region because of the difficulty of <span class="hlt">observations</span> by research vessels in this region, where territory problems between Japan and Russia is unsolved. Because of the scant of direct <span class="hlt">observation</span>, the Okhotsk Sea was still mysterious. In 2006 August, GPS radiosonde <span class="hlt">observation</span> was carried out by Russian research vessel Khromov in the Sea of Okhotsk by the cooperation between Japan and Russia, and strong SST <span class="hlt">gradient</span> of about 7 Celsius degrees/10km was <span class="hlt">observed</span> around the Kuril Islands. The purpose of this study is to demonstrate <span class="hlt">observational</span> finding of meso-scale atmospheric anticyclonic circulation influenced by the cold oceanic spot around the Kuril Island. The summaries of the <span class="hlt">observation</span> are as follows. Meso-scale atmospheric ageostrophic anticyclonic circulation in the atmospheric marine-boundary layer is <span class="hlt">observed</span> in and around the cold spot. A high air pressure area as compared with other surrounding areas is also located at the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730004177','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730004177"><span>Supplementary active stabilization of nonrigid gravity <span class="hlt">gradient</span> satellites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Keat, J. E.</p> <p>1972-01-01</p> <p>The use of active control for stability augmentation of passive gravity <span class="hlt">gradient</span> satellites is investigated. The reaction jet method of control is the main interest. Satellite nonrigidity is emphasized. The reduction in the Hamiltonian H is used as a control criteria. The <span class="hlt">velocities</span>, relative to local vertical, of the jets along their force axes are shown to be of fundamental significance. A basic control scheme which satisfies the H reduction criteria is developed. Each jet is fired when its <span class="hlt">velocity</span> becomes appropriately large. The jet is de-energized when <span class="hlt">velocity</span> reaches zero. Firing constraints to preclude orbit alteration may be needed. Control is continued until H has been minimized. This control policy is investigated using impulse and rectangular pulse models of the jet outputs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006A%26A...445..601G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006A%26A...445..601G"><span>Probing turbulence with infrared <span class="hlt">observations</span> in OMC1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gustafsson, M.; Field, D.; Lemaire, J. L.; Pijpers, F. P.</p> <p>2006-01-01</p> <p>A statistical analysis is presented of the turbulent <span class="hlt">velocity</span> structure in the Orion Molecular Cloud at scales ranging from 70 AU to 3×104 AU. Results are based on IR Fabry-Perot interferometric <span class="hlt">observations</span> of shock and photon-excited H2 in the K-band S(1) v=1{-}0 line at 2.121 μm and refer to the dynamical characteristics of warm perturbed gas. Data consist of a spatially resolved image with a measured <span class="hlt">velocity</span> for each resolution limited region (70 AU× 70 AU) in the image. The effect of removal of apparent large scale <span class="hlt">velocity</span> <span class="hlt">gradients</span> is discussed and the conclusion drawn that these apparent <span class="hlt">gradients</span> represent part of the turbulent cascade and should remain within the data. Using our full data set, <span class="hlt">observations</span> establish that the Larson size-linewidth relation is obeyed to the smallest scales studied here extending the range of validity of this relationship by nearly 2 orders of magnitude. The <span class="hlt">velocity</span> probability distribution function (PDF) is constructed showing extended exponential wings, providing evidence of intermittency, further supported by the skewness (third moment) and kurtosis (fourth moment) of the <span class="hlt">velocity</span> distribution. Variance and kurtosis of the PDF of <span class="hlt">velocity</span> differences are constructed as a function of lag. The variance shows an approximate power law dependence on lag, with exponent significantly lower than the Kolmogorov value, and with deviations below 2000 AU which are attributed to outflows and possibly disk structures associated with low mass star formation within OMC1. The kurtosis shows strong deviation from a Gaussian <span class="hlt">velocity</span> field, providing evidence of <span class="hlt">velocity</span> correlations at small lags. Results agree accurately with semi-empirical simulations in Eggers & Wang (1998). In addition, 170 individual H2 emitting clumps have been analysed with sizes between 500 and 2200 AU. These show considerable diversity with regard to PDFs and variance functions (related to second order structure functions) displaying a variety of shapes of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120013491','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120013491"><span>Inferring Lower Boundary Driving Conditions Using Vector Magnetic Field <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schuck, Peter W.; Linton, Mark; Leake, James; MacNeice, Peter; Allred, Joel</p> <p>2012-01-01</p> <p>Low-beta coronal MHD simulations of realistic CME events require the detailed specification of the magnetic fields, <span class="hlt">velocities</span>, densities, temperatures, etc., in the low corona. Presently, the most accurate estimates of solar vector magnetic fields are made in the high-beta photosphere. Several techniques have been developed that provide accurate estimates of the associated photospheric plasma <span class="hlt">velocities</span> such as the Differential Affine <span class="hlt">Velocity</span> Estimator for Vector Magnetograms and the Poloidal/Toroidal Decomposition. Nominally, these <span class="hlt">velocities</span> are consistent with the evolution of the radial magnetic field. To evolve the tangential magnetic field radial <span class="hlt">gradients</span> must be specified. In addition to estimating the photospheric vector magnetic and <span class="hlt">velocity</span> fields, a further challenge involves incorporating these fields into an MHD simulation. The simulation boundary must be driven, consistent with the numerical boundary equations, with the goal of accurately reproducing the <span class="hlt">observed</span> magnetic fields and estimated <span class="hlt">velocities</span> at some height within the simulation. Even if this goal is achieved, many unanswered questions remain. How can the photospheric magnetic fields and <span class="hlt">velocities</span> be propagated to the low corona through the transition region? At what cadence must we <span class="hlt">observe</span> the photosphere to realistically simulate the corona? How do we model the magnetic fields and plasma <span class="hlt">velocities</span> in the quiet Sun? How sensitive are the solutions to other unknowns that must be specified, such as the global solar magnetic field, and the photospheric temperature and density?</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013A%26A...559A..59B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013A%26A...559A..59B"><span>Chemical <span class="hlt">gradients</span> in the Milky Way from the RAVE data. I. Dwarf stars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boeche, C.; Siebert, A.; Piffl, T.; Just, A.; Steinmetz, M.; Sharma, S.; Kordopatis, G.; Gilmore, G.; Chiappini, C.; Williams, M.; Grebel, E. K.; Bland-Hawthorn, J.; Gibson, B. K.; Munari, U.; Siviero, A.; Bienaymé, O.; Navarro, J. F.; Parker, Q. A.; Reid, W.; Seabroke, G. M.; Watson, F. G.; Wyse, R. F. G.; Zwitter, T.</p> <p>2013-11-01</p> <p>-disc stars with a higher metallicity and Rg, generates a fictitious positive <span class="hlt">gradient</span> of the full sample. The flatter <span class="hlt">gradient</span> at larger Zmax found in the RAVE and the GCS samples may therefore be due to the superposition of thin- and thick-disc stars, which mimicks a flatter or positive <span class="hlt">gradient</span>. This does not exclude the possibility that the thick disc has no chemical <span class="hlt">gradient</span>. The discrepancies between the <span class="hlt">observational</span> samples and the mock sample can be reduced by i) decreasing the density; ii) decreasing the vertical <span class="hlt">velocity</span>; and iii) increasing the metallicity of the thick disc in the Besançon model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.S24B..07T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.S24B..07T"><span>P-wave <span class="hlt">Velocity</span> Structure in the Lowermost 600 km of the Mantle beneath Western Pacific Inferred from Travel Times and Amplitudes <span class="hlt">Observed</span> with NECESSArray</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tanaka, S.; Kawakatsu, H.; Chen, Y. J.; Ning, J.; Grand, S. P.; Niu, F.; Obayashi, M.; Miyakawa, K.; Idehara, K.; Tonegawa, T.; Iritani, R.; Necessarray Project Team</p> <p>2011-12-01</p> <p>NECESSArray is a large-scale broadband seismic array deployed in northeastern China. Although its primary aims are to reveal the fate of subducted Pacific plate and to address several tectonic issues, it is also useful as a large aperture array to look at deep Earth. Here, we examine P-wave travel times <span class="hlt">observed</span> with NECESSArray to determine P-wave <span class="hlt">velocity</span> structure in the lower mantle beneath Western Pacific. Relative travel times with respect to those predicted by PREM are measured on short period seismograms from 15 earthquakes occurred in Tonga, Fiji, and Kermadec regions since Sep. 2009 to April 2010, so far, by using adaptive stacking method [Rawlinson and Kennett, 2004]. The residuals are defined as fluctuations with respect to an average of the whole array for each event. Station correction is defined as a median value of the residuals at each station. After applying the station corrections and distance corrections for the surface focus, we synthesize all the residuals and finally obtain a characteristic residual variation as a function of epicentral distance from 80 to 95 degrees. The travel time residuals show an inverted V-pattern with the maximum delay of 0.2 s at 87 degrees compared from a reference level at 80 and 95 degrees. To simply interpret this pattern through Herglotz-Wiechert inversion, we assume that the <span class="hlt">velocity</span> structure above 600 km above the core-mantle boundary (CMB) is identical to PREM and find that the difference of the P-wave <span class="hlt">velocities</span> from those of PREM gradually increase with depth, and reach the maximum <span class="hlt">velocity</span> reduction of 0.15% and suddenly increase to those being identical to PREM at 270 km above the CMB. Thickness of a small <span class="hlt">velocity</span> <span class="hlt">gradient</span> layer at the base of the mantle is reduced to be 130 km instead of 150 km that is PREM's value. P-wave amplitudes are used as supplementary data. Station corrections for amplitude are inferred from 6 deep Fiji earthquakes in the distance range 75 to 90 degrees, in which focal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22521999-sma-observations-w3-oh-complex-physical-chemical-differentiation-between-w3-sub-w3-oh','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22521999-sma-observations-w3-oh-complex-physical-chemical-differentiation-between-w3-sub-w3-oh"><span>SMA <span class="hlt">OBSERVATIONS</span> OF THE W3(OH) COMPLEX: PHYSICAL AND CHEMICAL DIFFERENTIATION BETWEEN W3(H{sub 2}O) AND W3(OH)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Qin, Sheng-Li; Schilke, Peter; Sánchez-Monge, Álvaro</p> <p>2015-04-10</p> <p>We report on the Submillimeter Array (SMA) <span class="hlt">observations</span> of molecular lines at 270 GHz toward the W3(OH) and W3(H{sub 2}O) complex. Although previous <span class="hlt">observations</span> already resolved the W3(H{sub 2}O) into two or three sub-components, the physical and chemical properties of the two sources are not well constrained. Our SMA <span class="hlt">observations</span> clearly resolved the W3(OH) and W3(H{sub 2}O) continuum cores. Taking advantage of the line fitting tool XCLASS, we identified and modeled a rich molecular spectrum in this complex, including multiple CH{sub 3}CN and CH{sub 3}OH transitions in both cores. HDO, C{sub 2}H{sub 5}CN, O{sup 13}CS, and vibrationally excited lines ofmore » HCN, CH{sub 3}CN, and CH{sub 3}OCHO were only detected in W3(H{sub 2}O). We calculate gas temperatures and column densities for both cores. The results show that W3(H{sub 2}O) has higher gas temperatures and larger column densities than W3(OH) as previously <span class="hlt">observed</span>, suggesting physical and chemical differences between the two cores. We compare the molecular abundances in W3(H{sub 2}O) to those in the Sgr B2(N) hot core, the Orion KL hot core, and the Orion Compact Ridge, and discuss the chemical origin of specific species. An east–west <span class="hlt">velocity</span> <span class="hlt">gradient</span> is seen in W3(H{sub 2}O), and the extension is consistent with the bipolar outflow orientation traced by water masers and radio jets. A north–south <span class="hlt">velocity</span> <span class="hlt">gradient</span> across W3(OH) is also <span class="hlt">observed</span>. However, with current <span class="hlt">observations</span> we cannot be assured whether the <span class="hlt">velocity</span> <span class="hlt">gradients</span> are caused by rotation, outflow, or radial <span class="hlt">velocity</span> differences of the sub-components of W3(OH)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22441171','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22441171"><span><span class="hlt">Velocity</span> distributions in a micromixer measured by NMR imaging.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ahola, Susanna; Telkki, Ville-Veikko; Stapf, Siegfried</p> <p>2012-04-24</p> <p><span class="hlt">Velocity</span> distributions (so-called propagators) with two-dimensional spatial resolution inside a chemical micromixer were measured by pulsed-field-<span class="hlt">gradient</span> spin-echo (PGSE) nuclear magnetic resonance (NMR). A surface coil matching the volume of interest was built to enhance the signal-to-noise ratio. This enabled the acquisition of <span class="hlt">velocity</span> maps with a very high spatial resolution of 29 μm × 39 μm. The measured propagators are compared with theoretical distributions and a good agreement is found. The results show that the propagator data provide much richer information about flow behaviour than conventional NMR <span class="hlt">velocity</span> imaging and the information is essential for understanding the performance of a micromixer. It reveals, for example, deviations in the shape and size of the channel structures and multicomponent flow <span class="hlt">velocity</span> distribution of overlapping channels. Propagator data efficiently compensate lost information caused by insufficient 3D resolution in conventional <span class="hlt">velocity</span> imaging.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24710761','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24710761"><span>An optimized design to reduce eddy current sensitivity in <span class="hlt">velocity</span>-selective arterial spin labeling using symmetric BIR-8 pulses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Guo, Jia; Meakin, James A; Jezzard, Peter; Wong, Eric C</p> <p>2015-03-01</p> <p><span class="hlt">Velocity</span>-selective arterial spin labeling (VSASL) tags arterial blood on a <span class="hlt">velocity</span>-selective (VS) basis and eliminates the tagging/imaging gap and associated transit delay sensitivity <span class="hlt">observed</span> in other ASL tagging methods. However, the flow-weighting <span class="hlt">gradient</span> pulses in VS tag preparation can generate eddy currents (ECs), which may erroneously tag the static tissue and create artificial perfusion signal, compromising the accuracy of perfusion quantification. A novel VS preparation design is presented using an eight-segment B1 insensitive rotation with symmetric radio frequency and <span class="hlt">gradient</span> layouts (sym-BIR-8), combined with delays after <span class="hlt">gradient</span> pulses to optimally reduce ECs of a wide range of time constants while maintaining B0 and B1 insensitivity. Bloch simulation, phantom, and in vivo experiments were carried out to determine robustness of the new and existing pulse designs to ECs, B0 , and B1 inhomogeneity. VSASL with reduced EC sensitivity across a wide range of EC time constants was achieved with the proposed sym-BIR-8 design, and the accuracy of cerebral blood flow measurement was improved. The sym-BIR-8 design performed the most robustly among the existing VS tagging designs, and should benefit studies using VS preparation with improved accuracy and reliability. © 2014 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120016411','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120016411"><span>Calculating the <span class="hlt">Velocity</span> in the Moss</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Womebarger, Amy R.; Tripathi, Durgesh; Mason, Helen</p> <p>2011-01-01</p> <p>The <span class="hlt">velocity</span> of the warm (1 MK) plasma in the footpoint of the hot coronal loops (commonly called moss) could help discriminate between different heating frequencies in the active region core. Strong <span class="hlt">velocities</span> would indicated low-frequency heating, while <span class="hlt">velocities</span> close to zero would indicate high-frequency heating. Previous results have found disparaging <span class="hlt">observations</span>, with both strong <span class="hlt">velocities</span> and <span class="hlt">velocities</span> close to zero reported. Previous results are based on <span class="hlt">observations</span> from Hinode/EIS. The wavelength arrays for EIS spectra are typically calculated by assuming quiet Sun <span class="hlt">velocities</span> are zero. In this poster, we determine the <span class="hlt">velocity</span> in the moss using <span class="hlt">observations</span> with SoHO/SUMER. We rely on neutral or singly ionized spectral lines to determine accurately the wavelength array associated with the spectra. SUMER scanned the active region twice, so we also report the stability of the <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A43C2468M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A43C2468M"><span>Constraints on Southern Ocean CO2 Fluxes and Seasonality from Atmospheric Vertical <span class="hlt">Gradients</span> <span class="hlt">Observed</span> on Multiple Airborne Campaigns</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McKain, K.; Sweeney, C.; Stephens, B. B.; Long, M. C.; Jacobson, A. R.; Basu, S.; Chatterjee, A.; Weir, B.; Wofsy, S. C.; Atlas, E. L.; Blake, D. R.; Montzka, S. A.; Stern, R.</p> <p>2017-12-01</p> <p>The Southern Ocean plays an important role in the global carbon cycle and climate system, but net CO2 flux into the Southern Ocean is difficult to measure and model because it results from large opposing and seasonally-varying fluxes due to thermal forcing, biological uptake, and deep-water mixing. We present an analysis to constrain the seasonal cycle of net CO2 exchange with the Southern Ocean, and the magnitude of summer uptake, using the vertical <span class="hlt">gradients</span> in atmospheric CO2 <span class="hlt">observed</span> during three aircraft campaigns in the southern polar region. The O2/N2 Ratio and CO2 Airborne Southern Ocean Study (ORCAS) was an airborne campaign that intensively sampled the atmosphere at 0-13 km altitude and 45-75 degrees south latitude in the austral summer (January-February) of 2016. The global airborne campaigns, the HIAPER Pole-to-Pole <span class="hlt">Observations</span> (HIPPO) study and the Atmospheric Tomography Mission (ATom), provide additional measurements over the Southern Ocean from other seasons and multiple years (2009-2011, 2016-2017). Derivation of fluxes from measured vertical <span class="hlt">gradients</span> requires robust estimates of the residence time of air in the polar tropospheric domain, and of the contribution of long-range transport from northern latitudes outside the domain to the CO2 <span class="hlt">gradient</span>. We use diverse independent approaches to estimate both terms, including simulations using multiple transport and flux models, and <span class="hlt">observed</span> <span class="hlt">gradients</span> of shorter-lived tracers with specific sources regions and well-known loss processes. This study demonstrates the utility of aircraft profile measurements for constraining large-scale air-sea fluxes for the Southern Ocean, in contrast to those derived from the extrapolation of sparse ocean and atmospheric measurements and uncertain flux parameterizations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800009140','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800009140"><span>First results of a study on turbulent boundary layers in oscillating flow with a mean adverse pressure <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Houdeville, R.; Cousteix, J.</p> <p>1979-01-01</p> <p>The development of a turbulent unsteady boundary layer with a mean pressure <span class="hlt">gradient</span> strong enough to induce separation, in order to complete the extend results obtained for the flat plate configuration is presented. The longitudinal component of the <span class="hlt">velocity</span> is measured using constant temperature hot wire anemometer. The region where negative <span class="hlt">velocities</span> exist is investigated with a laser Doppler velocimeter system with BRAGG cells. The boundary layer responds by forced pulsation to the perturbation of potential flow. The unsteady effects <span class="hlt">observed</span> are very important. The average location of the zero skin friction point moves periodically at the perturbation frequency. Average <span class="hlt">velocity</span> profiles from different instants in the cycle are compared. The existence of a logarithmic region enables a simple calculation of the maximum phase shift of the <span class="hlt">velocity</span> in the boundary layer. An attempt of calculation by an integral method of boundary layer development is presented, up to the point where reverse flow starts appearing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JGRD..119.9707M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JGRD..119.9707M"><span>New statistical analysis of the horizontal phase <span class="hlt">velocity</span> distribution of gravity waves <span class="hlt">observed</span> by airglow imaging</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Matsuda, Takashi S.; Nakamura, Takuji; Ejiri, Mitsumu K.; Tsutsumi, Masaki; Shiokawa, Kazuo</p> <p>2014-08-01</p> <p>We have developed a new analysis method for obtaining the power spectrum in the horizontal phase <span class="hlt">velocity</span> domain from airglow intensity image data to study atmospheric gravity waves. This method can deal with extensive amounts of imaging data obtained on different years and at various <span class="hlt">observation</span> sites without bias caused by different event extraction criteria for the person processing the data. The new method was applied to sodium airglow data obtained in 2011 at Syowa Station (69°S, 40°E), Antarctica. The results were compared with those obtained from a conventional event analysis in which the phase fronts were traced manually in order to estimate horizontal characteristics, such as wavelengths, phase <span class="hlt">velocities</span>, and wave periods. The horizontal phase <span class="hlt">velocity</span> of each wave event in the airglow images corresponded closely to a peak in the spectrum. The statistical results of spectral analysis showed an eastward offset of the horizontal phase <span class="hlt">velocity</span> distribution. This could be interpreted as the existence of wave sources around the stratospheric eastward jet. Similar zonal anisotropy was also seen in the horizontal phase <span class="hlt">velocity</span> distribution of the gravity waves by the event analysis. Both methods produce similar statistical results about directionality of atmospheric gravity waves. Galactic contamination of the spectrum was examined by calculating the apparent <span class="hlt">velocity</span> of the stars and found to be limited for phase speeds lower than 30 m/s. In conclusion, our new method is suitable for deriving the horizontal phase <span class="hlt">velocity</span> characteristics of atmospheric gravity waves from an extensive amount of imaging data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997DPS....29.3723R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997DPS....29.3723R"><span>Fabry-Perot <span class="hlt">Observations</span> of Comet Hale-Bopp H_2O(+) <span class="hlt">Velocity</span> Fields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roesler, F. L.; Klinglesmith, D. A., III; Scherb, F.; Mierkiewicz, E. J.; Oliversen, R. J.</p> <p>1997-07-01</p> <p>We have obtained Doppler-sliced images of H_2O(+) emission from Comet Hale-Bopp, using a 15-cm, dual-etalon, Fabry-Perot/CCD imaging spectrometer at the McMath-Pierce 0.8-meter west auxiliary telescope of the National Solar Observatory on Kitt Peak. The 6-arcmin field of view was centered on the comet nucleus, and the spectral resolution was 0.4 Angstroms (20km/sec). The <span class="hlt">observations</span> consisted of ``data cubes,'' i.e., a sequence of images of the 6158 Angstroms emission doublet at <span class="hlt">velocity</span> steps of 12.5 or 25km/sec, covering a range from -75km/sec to +75km/sec in the comet reference frame. We were able to follow the comet for 1 to 1(1/_2) hours each clear night. We obtained useable data cubes on at least ten nights between February 25 and April 16. These data are being examined to investigate the comet-solar wind interaction. We will present both still images and time-lapse movies showing sequences of ion <span class="hlt">velocities</span> and accelerations on the plane of the sky.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ASPC..478..145S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ASPC..478..145S"><span>Differences of the Solar Magnetic Activity Signature in <span class="hlt">Velocity</span> and Intensity Helioseismic <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Salabert, D.; García, R. A.; Jiménez, A.</p> <p>2013-12-01</p> <p>The high-quality, full-disk helioseismic <span class="hlt">observations</span> continuously collected by the spectrophotometer GOLF and the three photometers VIRGO/SPMs onboard the SoHO spacecraft for 17 years now (since April 11, 1996, apart from the SoHO “vacations”) are absolutely unique for the study of the interior of the Sun and its variability with magnetic activity. Here, we look at the differences in the low-degree oscillation p-mode frequencies between radial <span class="hlt">velocity</span> and intensity measurements taking into account all the known features of the p-mode profiles (e.g., the opposite peak asymmetry), and of the power spectrum (e.g., the presence of the higher degrees ℓ = 4 and 5 in the signal). We show that the intensity frequencies are higher than the <span class="hlt">velocity</span> frequencies during the solar cycle with a clear temporal dependence. The response between the individual angular degrees is also different. Time delays are <span class="hlt">observed</span> between the temporal variations in GOLF and VIRGO frequencies. Such analysis is important in order to put new constraints and to better understand the mechanisms responsible for the temporal variations of the oscillation frequencies with the solar magnetic activity as well as their height dependences in the solar atmosphere. It is also important for the study of the stellar magnetic activity using asteroseismic data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4473749','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4473749"><span>Mapping conduction <span class="hlt">velocity</span> of early embryonic hearts with a robust fitting algorithm</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gu, Shi; Wang, Yves T; Ma, Pei; Werdich, Andreas A; Rollins, Andrew M; Jenkins, Michael W</p> <p>2015-01-01</p> <p>Cardiac conduction maturation is an important and integral component of heart development. Optical mapping with voltage-sensitive dyes allows sensitive measurements of electrophysiological signals over the entire heart. However, accurate measurements of conduction <span class="hlt">velocity</span> during early cardiac development is typically hindered by low signal-to-noise ratio (SNR) measurements of action potentials. Here, we present a novel image processing approach based on least squares optimizations, which enables high-resolution, low-noise conduction <span class="hlt">velocity</span> mapping of smaller tubular hearts. First, the action potential trace measured at each pixel is fit to a curve consisting of two cumulative normal distribution functions. Then, the activation time at each pixel is determined based on the fit, and the spatial <span class="hlt">gradient</span> of activation time is determined with a two-dimensional (2D) linear fit over a square-shaped window. The size of the window is adaptively enlarged until the <span class="hlt">gradients</span> can be determined within a preset precision. Finally, the conduction <span class="hlt">velocity</span> is calculated based on the activation time <span class="hlt">gradient</span>, and further corrected for three-dimensional (3D) geometry that can be obtained by optical coherence tomography (OCT). We validated the approach using published activation potential traces based on computer simulations. We further validated the method by adding artificially generated noise to the signal to simulate various SNR conditions using a curved simulated image (digital phantom) that resembles a tubular heart. This method proved to be robust, even at very low SNR conditions (SNR = 2-5). We also established an empirical equation to estimate the maximum conduction <span class="hlt">velocity</span> that can be accurately measured under different conditions (e.g. sampling rate, SNR, and pixel size). Finally, we demonstrated high-resolution conduction <span class="hlt">velocity</span> maps of the quail embryonic heart at a looping stage of development. PMID:26114034</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1982PhDT.........2Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1982PhDT.........2Y"><span>Thermal-<span class="hlt">gradient</span> migration of brine inclusions in salt crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yagnik, S. K.</p> <p>1982-09-01</p> <p>High level nuclear waste disposal in a geologic repository was proposed. Natural salt deposits which are considered contain a small volume fraction of water in the form of brine inclusions distributed throughout the salt. Radioactive decay heating of the nuclear wastes will impose a temperature <span class="hlt">gradient</span> on the surrounding salt which mobilizes the brine inclusions. Inclusions filled completely with brine migrate up the temperature <span class="hlt">gradient</span> and eventually accumulate brine near the buried waste forms. The brine may slowly corrode or degrade the waste forms which is undesirable. In this work, thermal <span class="hlt">gradient</span> migration of both all liquid and gas liquid inclusions was experimentally studied in synthetic single crystals of NaCl and KCl using a hot stage attachment to an optical microscope which was capable of imposing temperature <span class="hlt">gradients</span> and axial compressive loads on the crystals. The migration <span class="hlt">velocities</span> of the inclusion shape and size are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19750009639','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19750009639"><span>Optimization of structures to satisfy a flutter <span class="hlt">velocity</span> constraint by use of quadratic equation fitting. M.S. Thesis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Motiwalla, S. K.</p> <p>1973-01-01</p> <p>Using the first and the second derivative of flutter <span class="hlt">velocity</span> with respect to the parameters, the <span class="hlt">velocity</span> hypersurface is made quadratic. This greatly simplifies the numerical procedure developed for determining the values of the design parameters such that a specified flutter <span class="hlt">velocity</span> constraint is satisfied and the total structural mass is near a relative minimum. A search procedure is presented utilizing two <span class="hlt">gradient</span> search methods and a <span class="hlt">gradient</span> projection method. The procedure is applied to the design of a box beam, using finite-element representation. The results indicate that the procedure developed yields substantial design improvement satisfying the specified constraint and does converge to near a local optimum.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.G11A0857E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.G11A0857E"><span>Detection of spatio-temporal change of ocean acoustic <span class="hlt">velocity</span> for <span class="hlt">observing</span> seafloor crustal deformation applying seismological methods</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eto, S.; Nagai, S.; Tadokoro, K.</p> <p>2011-12-01</p> <p>Our group has developed a system for <span class="hlt">observing</span> seafloor crustal deformation with a combination of acoustic ranging and kinematic GPS positioning techniques. One of the effective factors to reduce estimation error of submarine benchmark in our system is modeling variation of ocean acoustic <span class="hlt">velocity</span>. We estimated various 1-dimensional <span class="hlt">velocity</span> models with depth under some constraints, because it is difficult to estimate 3-dimensional acoustic <span class="hlt">velocity</span> structure including temporal change due to our simple acquisition procedure of acoustic ranging data. We, then, applied the joint hypocenter determination method in seismology [Kissling et al., 1994] to acoustic ranging data. We assume two conditions as constraints in inversion procedure as follows: 1) fixed acoustic <span class="hlt">velocity</span> in deeper part because it is usually stable both in space and time, 2) each inverted <span class="hlt">velocity</span> model should be decreased with depth. The following two remarkable spatio-temporal changes of acoustic <span class="hlt">velocity</span> 1) variations of travel-time residuals at the same points within short time and 2) larger differences between residuals at the neighboring points, which are one's of travel-time from different benchmarks. The First results cannot be explained only by the effect of atmospheric condition change including heating by sunlight. To verify the residual variations mentioned as the second result, we have performed forward modeling of acoustic ranging data with <span class="hlt">velocity</span> models added <span class="hlt">velocity</span> anomalies. We calculate travel time by a pseudo-bending ray tracing method [Um and Thurber, 1987] to examine effects of <span class="hlt">velocity</span> anomaly on the travel-time differences. Comparison between these residuals and travel-time difference in forward modeling, <span class="hlt">velocity</span> anomaly bodies in shallower depth can make these anomalous residuals, which may indicate moving water bodies. We need to apply an acoustic <span class="hlt">velocity</span> structure model with <span class="hlt">velocity</span> anomaly(s) in acoustic ranging data analysis and/or to develop a new system with a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MPLB...3240003D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MPLB...3240003D"><span>Stability of boundary layer flow based on energy <span class="hlt">gradient</span> theory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dou, Hua-Shu; Xu, Wenqian; Khoo, Boo Cheong</p> <p>2018-05-01</p> <p>The flow of the laminar boundary layer on a flat plate is studied with the simulation of Navier-Stokes equations. The mechanisms of flow instability at external edge of the boundary layer and near the wall are analyzed using the energy <span class="hlt">gradient</span> theory. The simulation results show that there is an overshoot on the <span class="hlt">velocity</span> profile at the external edge of the boundary layer. At this overshoot, the energy <span class="hlt">gradient</span> function is very large which results in instability according to the energy <span class="hlt">gradient</span> theory. It is found that the transverse <span class="hlt">gradient</span> of the total mechanical energy is responsible for the instability at the external edge of the boundary layer, which induces the entrainment of external flow into the boundary layer. Within the boundary layer, there is a maximum of the energy <span class="hlt">gradient</span> function near the wall, which leads to intensive flow instability near the wall and contributes to the generation of turbulence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhRvE..69a1201S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhRvE..69a1201S"><span>Thermophoresis of dissolved molecules and polymers: Consideration of the temperature-induced macroscopic pressure <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Semenov, Semen; Schimpf, Martin</p> <p>2004-01-01</p> <p>The movement of molecules and homopolymer chains dissolved in a nonelectrolyte solvent in response to a temperature <span class="hlt">gradient</span> is considered a consequence of temperature-induced pressure <span class="hlt">gradients</span> in the solvent layer surrounding the solute molecules. Local pressure <span class="hlt">gradients</span> are produced by nonuniform London van der Waals interactions, established by <span class="hlt">gradients</span> in the concentration (density) of solvent molecules. The density <span class="hlt">gradient</span> is produced by variations in solvent thermal expansion within the nonuniform temperature field. The resulting expression for the <span class="hlt">velocity</span> of the solute contains the Hamaker constants for solute-solvent and solute-solute interactions, the radius of the solute molecule, and the viscosity and cubic coefficient of thermal expansion of the solvent. In this paper we consider an additional force that arises from directional asymmetry in the interaction between solvent molecules. In a closed cell, the resulting macroscopic pressure <span class="hlt">gradient</span> gives rise to a volume force that affects the motion of dissolved solutes. An expression for this macroscopic pressure <span class="hlt">gradient</span> is derived and the resulting force is incorporated into the expression for the solute <span class="hlt">velocity</span>. The expression is used to calculate thermodiffusion coefficients for polystyrene in several organic solvents. When these values are compared to those measured in the laboratory, the consistency is better than that found in previous reports, which did not consider the macroscopic pressure <span class="hlt">gradient</span> that arises in a closed thermodiffusion cell. The model also allows for the movement of solute in either direction, depending on the relative values of the solvent and solute Hamaker constants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999JGR...104.4783Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999JGR...104.4783Z"><span>Upper mantle <span class="hlt">velocity</span> structure beneath southern Africa from modeling regional seismic data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Ming; Langston, Charles A.; Nyblade, Andrew A.; Owens, Thomas J.</p> <p>1999-03-01</p> <p>The upper mantle seismic <span class="hlt">velocity</span> structure beneath southern Africa is investigated using travel time and waveform data which come from a large mine tremor in South Africa (mb 5.6) recorded by the Tanzania broadband seismic experiment and by several stations in southern Africa. The waveform data show upper mantle triplications for both the 410- and 670-km discontinuities between distances of 2100 and 3000 km. Auxiliary travel time data along similar profiles obtained from other moderate events are also used. P wave travel times are inverted for <span class="hlt">velocity</span> structure down to ˜800-km depth using the Wiechert-Herglotz technique, and the resulting model is evaluated by perturbing it at three depth intervals and then testing the perturbed model against the travel time and waveform data. The results indicate a typical upper mantle P wave <span class="hlt">velocity</span> structure for a shield. P wave <span class="hlt">velocities</span> from the top of the mantle down to 300-km depth are as much as 3% higher than the global average and are slightly slower than the global average between 300- and 420-km depth. Little evidence is found for a pronounced low-<span class="hlt">velocity</span> zone in the upper mantle. A high-<span class="hlt">velocity</span> <span class="hlt">gradient</span> zone is required above the 410-km discontinuity, but both sharp and smooth 410-km discontinuities are permitted by the data. The 670-km discontinuity is characterized by high-<span class="hlt">velocity</span> <span class="hlt">gradients</span> over a depth range of ˜80 km around 660-km depth. Limited S wave travel time data suggest fast S wave <span class="hlt">velocities</span> above ˜150-km depth. These results suggest that the bouyant support for the African superswell does not reside at shallow depths in the upper mantle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011APS..DFDR21005K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011APS..DFDR21005K"><span>The effect of non-zero radial <span class="hlt">velocity</span> on the impulse and circulation of starting jets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krieg, Michael; Mohseni, Kamran</p> <p>2011-11-01</p> <p>Vortex ring formation dynamics are generally studied using two basic types of vortex generators. Piston cylinder vortex generators eject fluid through a long tube which ensures a purely axial jet; whereas, vortex ring generators which expel fluid through a flat plate with a circular orifice produce 2-D jets (non-zero radial <span class="hlt">velocity</span>). At the nozzle exit plane of the orifice type vortex generator the radial component of <span class="hlt">velocity</span> is linearly proportional to the radial distance from the axis of symmetry, reaching a maximum at the edge of the orifice with a magnitude around 10 % of the piston <span class="hlt">velocity</span> (the ratio of the volume flux and the nozzle area). As the jet advances downstream the radial <span class="hlt">velocity</span> quickly dissipates, and becomes purely axial less than a diameter away from the nozzle exit plane. The radial <span class="hlt">velocity</span> <span class="hlt">gradient</span> in the axial direction plays a key role in the rate at which circulation and impulse are ejected from the vortex generator. Though the radial component of <span class="hlt">velocity</span> is small compared to the axial <span class="hlt">velocity</span>, it has a significant effect on both the circulation and impulse of the starting jet because of this <span class="hlt">gradient</span>. The extent of circulation and impulse enhancement is investigated through experimental DPIV data showing that the orifice device produces nearly double both circulation and energy (with identical piston <span class="hlt">velocity</span> and stroke ratios).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19750057152&hterms=1088&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3D%2526%25231088','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19750057152&hterms=1088&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3D%2526%25231088"><span>Cosmic ray intensity <span class="hlt">gradients</span> in the solar system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mckibben, R. B.</p> <p>1975-01-01</p> <p>Recent progress in the determination of cosmic-ray intensity <span class="hlt">gradients</span> is reviewed. Direct satellite measurements of the integral <span class="hlt">gradient</span> are described together with various types of indirect measurements, including measurements of the Ar-37/Ar-39 ratio in samples from the Lost City meteorite, studies of anisotropies in neutron-monitor counting rates, and analysis of the sidereal diurnal anisotropy <span class="hlt">observed</span> at a single point on earth. Nucleonic radial <span class="hlt">gradients</span> and electron <span class="hlt">gradients</span> measured by satellites in differential energy windows are discussed, and theoretical studies of the physical processes involved in these <span class="hlt">gradients</span> are summarized. <span class="hlt">Observations</span> of intensity <span class="hlt">gradients</span> in heliographic latitude are reported.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003NucFu..43..228G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003NucFu..43..228G"><span>Theory and <span class="hlt">observations</span> of high frequency Alfvén eigenmodes in low aspect ratio plasmas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gorelenkov, N. N.; Fredrickson, E.; Belova, E.; Cheng, C. Z.; Gates, D.; Kaye, S.; White, R.</p> <p>2003-04-01</p> <p>New <span class="hlt">observations</span> of sub-cyclotron frequency instability in low aspect ratio plasmas in national spherical torus experiments are reported. The frequencies of <span class="hlt">observed</span> instabilities correlate with the characteristic Alfvén <span class="hlt">velocity</span> of the plasma. A theory of localized compressional Alfvén eigenmodes (CAE) and global shear Alfvén eigenmodes (GAE) in low aspect ratio plasmas is presented to explain the <span class="hlt">observed</span> high frequency instabilities. CAEs/GAEs are driven by the <span class="hlt">velocity</span> space <span class="hlt">gradient</span> of energetic super-Alfvénic beam ions via Doppler shifted cyclotron resonances. One of the main damping mechanisms of GAEs, the continuum damping, is treated perturbatively within the framework of ideal MHD. Properties of these cyclotron instability ions are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930092289','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930092289"><span>Similar solutions for the compressible laminar boundary layer with heat transfer and pressure <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cohen, Clarence B; Reshotko, Eli</p> <p>1956-01-01</p> <p>Stewartson's transformation is applied to the laminar compressible boundary-layer equations and the requirement of similarity is introduced, resulting in a set of ordinary nonlinear differential equations previously quoted by Stewartson, but unsolved. The requirements of the system are Prandtl number of 1.0, linear viscosity-temperature relation across the boundary layer, an isothermal surface, and the particular distributions of free-stream <span class="hlt">velocity</span> consistent with similar solutions. This system admits axial pressure <span class="hlt">gradients</span> of arbitrary magnitude, heat flux normal to the surface, and arbitrary Mach numbers. The system of differential equations is transformed to integral system, with the <span class="hlt">velocity</span> ratio as the independent variable. For this system, solutions are found by digital computation for pressure <span class="hlt">gradients</span> varying from that causing separation to the infinitely favorable <span class="hlt">gradient</span> and for wall temperatures from absolute zero to twice the free-stream stagnation temperature. Some solutions for separated flows are also presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012CSR....48...87W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012CSR....48...87W"><span>Tidal asymmetries of <span class="hlt">velocity</span> and stratification over a bathymetric depression in a tropical inlet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Waterhouse, Amy F.; Valle-Levinson, Arnoldo; Morales Pérez, Rubén A.</p> <p>2012-10-01</p> <p><span class="hlt">Observations</span> of current <span class="hlt">velocity</span>, sea surface elevation and vertical profiles of density were obtained in a tropical inlet to determine the effect of a bathymetric depression (hollow) on the tidal flows. Surveys measuring <span class="hlt">velocity</span> profiles were conducted over a diurnal tidal cycle with mixed spring tides during dry and wet seasons. Depth-averaged tidal <span class="hlt">velocities</span> during ebb and flood tides behaved according to Bernoulli dynamics, as expected. The dynamic balance of depth-averaged quantities in the along-channel direction was governed by along-channel advection and pressure <span class="hlt">gradients</span> with baroclinic pressure <span class="hlt">gradients</span> only being important during the wet season. The vertical structure of the along-channel flow during flood tides exhibited a mid-depth maximum with lateral shear enhanced during the dry season as a result of decreased vertical stratification. During ebb tides, along-channel <span class="hlt">velocities</span> in the vicinity of the hollow were vertically sheared with a weak return flow at depth due to choking of the flow on the seaward slope of the hollow. The potential energy anomaly, a measure of the amount of energy required to fully mix the water column, showed two peaks in stratification associated with ebb tide and a third peak occurring at the beginning of flood. After the first mid-ebb peak in stratification, ebb flows were constricted on the seaward slope of the hollow resulting in a bottom return flow. The sinking of surface waters and enhanced mixing on the seaward slope of the hollow reduced the potential energy anomaly after maximum ebb. The third peak in stratification during early flood occurred as a result of denser water entering the inlet at mid-depth. This dense water mixed with ambient deep waters increasing the stratification. Lateral shear in the along-channel flow across the hollow allowed trapping of less dense water in the surface layers further increasing stratification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPhCS.708a2012B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPhCS.708a2012B"><span>Large-eddy simulations of adverse pressure <span class="hlt">gradient</span> turbulent boundary layers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bobke, Alexandra; Vinuesa, Ricardo; Örlü, Ramis; Schlatter, Philipp</p> <p>2016-04-01</p> <p>Adverse pressure-<span class="hlt">gradient</span> (APG) turbulent boundary layers (TBL) are studied by performing well-resolved large-eddy simulations. The pressure <span class="hlt">gradient</span> is imposed by defining the free-stream <span class="hlt">velocity</span> distribution with the description of a power law. Different inflow conditions, box sizes and upper boundary conditions are tested in order to determine the final set-up. The statistics of turbulent boundary layers with two different power-law coefficients and thus magnitudes of adverse pressure <span class="hlt">gradients</span> are then compared to zero pressure-<span class="hlt">gradient</span> (ZPG) data. The effect of the APG on TBLs is manifested in the mean flow through a much more prominent wake region and in the Reynolds stresses through the existence of an outer peak. The pre-multiplied energy budgets show that more energy is transported from the near-wall region to farther away from the wall.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18..497L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18..497L"><span><span class="hlt">Velocity</span> changes at Volcán de Colima: Seismic and Experimental <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lamb, Oliver; Lavallée, Yan; De Angelis, Silvio; Varley, Nick; Reyes-Dávila, Gabriel; Arámbula-Mendoza, Raúl; Hornby, Adrian; Wall, Richard; Kendrick, Jackie</p> <p>2016-04-01</p> <p>Immediately prior to dome-building eruptions, volcano-seismic swarms are a direct consequence of strain localisation in the ascending magma. A deformation mechanism map of magma subjected to strain localisation will help develop accurate numerical models, which, coupled to an understanding of the mechanics driving monitored geophysical signals prior to lava eruption, will enhance forecasts. Here we present how seismic data from Volcán de Colima, Mexico, is combined with experimental work to give insights into fracturing in and around magma. Volcán de Colima is a dome-forming volcano that has been almost-continuously erupting since November 1998. We use coda-wave interferometry to quantify small changes in seismic <span class="hlt">velocity</span> structure between pairs of similar earthquakes, employing waveforms from clusters of repeating earthquakes. The changes in all pairs of events were then used together to create a continuous function of <span class="hlt">velocity</span> change at all stations within 7 km of the volcano from October to December 1998. We complement our seismic data with acoustic emission data from tensional experiments using samples collected at Volcán de Colima. Decreases in <span class="hlt">velocity</span> and frequency reflect changes in the sample properties prior to failure. By comparing experimental and seismic <span class="hlt">observations</span>, we may place constraints on the conditions of the natural seismogenic processes. Using a combination of field and experimental data promises a greater understanding of the processes affecting the rise of magma during an eruption. This will help with the challenge of forecasting and hazard mitigation during dome-forming eruptions worldwide.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ITNS...64..829V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ITNS...64..829V"><span>The Effect of Non-Uniform Temperature and <span class="hlt">Velocity</span> Fields on Long Range Ultrasonic Measurement Systems in MYRRHA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Van De Wyer, Nicolas; Schram, Christophe; Van Dyck, Dries; Dierckx, Marc</p> <p>2017-02-01</p> <p>SCK·CEN, the Belgian Nuclear Research Center, is developing MYRRHA, a generation IV liquid metal cooled nuclear research reactor. As the liquid metal coolant is opaque to light, normal visual feedback during fuel manipulations is not available and must therefore be replaced by a system that is not hindered by the opacity of the coolant. In this respect ultrasonic based instrumentation is under development at SCK·CEN to provide feedback during operations under liquid metal. One of the tasks that will be tackled using ultrasound is the detection and localization of a potentially lost fuel assembly. The development of this localization tool is detailed in this paper. In this application, the distance between ultrasonic sensor and target may be as large as 2.5m. At these distances, non uniform <span class="hlt">velocity</span> and temperature fields in the liquid metal potentially influence the propagation of the ultrasonic signals, affecting the performance of the ultrasonic systems. In this paper, we investigate how relevant temperature and <span class="hlt">velocity</span> <span class="hlt">gradients</span> inside the liquid metal influence the propagation of ultrasonic waves. The effect of temperature and <span class="hlt">velocity</span> <span class="hlt">gradients</span> are simulated by means of a newly developed numerical raytracing model. The performance of the model is validated by dedicated water experiments. The setup is capable of creating <span class="hlt">velocity</span> and temperature <span class="hlt">gradients</span> representative for MYRRHA conditions. Once validated in water, the same model is used to make predictions for the effect of <span class="hlt">gradients</span> in the MYRRHA liquid metal environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28534494','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28534494"><span>Airborne <span class="hlt">observations</span> reveal elevational <span class="hlt">gradient</span> in tropical forest isoprene emissions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gu, Dasa; Guenther, Alex B; Shilling, John E; Yu, Haofei; Huang, Maoyi; Zhao, Chun; Yang, Qing; Martin, Scot T; Artaxo, Paulo; Kim, Saewung; Seco, Roger; Stavrakou, Trissevgeni; Longo, Karla M; Tóta, Julio; de Souza, Rodrigo Augusto Ferreira; Vega, Oscar; Liu, Ying; Shrivastava, Manish; Alves, Eliane G; Santos, Fernando C; Leng, Guoyong; Hu, Zhiyuan</p> <p>2017-05-23</p> <p>Isoprene dominates global non-methane volatile organic compound emissions, and impacts tropospheric chemistry by influencing oxidants and aerosols. Isoprene emission rates vary over several orders of magnitude for different plants, and characterizing this immense biological chemodiversity is a challenge for estimating isoprene emission from tropical forests. Here we present the isoprene emission estimates from aircraft eddy covariance measurements over the Amazonian forest. We report isoprene emission rates that are three times higher than satellite top-down estimates and 35% higher than model predictions. The results reveal strong correlations between <span class="hlt">observed</span> isoprene emission rates and terrain elevations, which are confirmed by similar correlations between satellite-derived isoprene emissions and terrain elevations. We propose that the elevational <span class="hlt">gradient</span> in the Amazonian forest isoprene emission capacity is determined by plant species distributions and can substantially explain isoprene emission variability in tropical forests, and use a model to demonstrate the resulting impacts on regional air quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ApJ...713.1376F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ApJ...713.1376F"><span>Damping of Magnetohydrodynamic Turbulence in Partially Ionized Gas and the <span class="hlt">Observed</span> Difference of <span class="hlt">Velocities</span> of Neutrals and Ions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Falceta-Gonçalves, D.; Lazarian, A.; Houde, M.</p> <p>2010-04-01</p> <p>Theoretical and <span class="hlt">observational</span> studies on the turbulence of the interstellar medium developed fast in the past decades. The theory of supersonic magnetized turbulence, as well as the understanding of the projection effects of <span class="hlt">observed</span> quantities, is still in progress. In this work, we explore the characterization of the turbulent cascade and its damping from <span class="hlt">observational</span> spectral line profiles. We address the difference of ion and neutral <span class="hlt">velocities</span> by clarifying the nature of the turbulence damping in the partially ionized. We provide theoretical arguments in favor of the explanation of the larger Doppler broadening of lines arising from neutral species compared to ions as arising from the turbulence damping of ions at larger scales. Also, we compute a number of MHD numerical simulations for different turbulent regimes and explicit turbulent damping, and compare both the three-dimensional distributions of <span class="hlt">velocity</span> and the synthetic line profile distributions. From the numerical simulations, we place constraints on the precision with which one can measure the three-dimensional dispersion depending on the turbulence sonic Mach number. We show that no universal correspondence between the three-dimensional <span class="hlt">velocity</span> dispersions measured in the turbulent volume and minima of the two-dimensional <span class="hlt">velocity</span> dispersions available through <span class="hlt">observations</span> exist. For instance, for subsonic turbulence the correspondence is poor at scales much smaller than the turbulence injection scale, while for supersonic turbulence the correspondence is poor for the scales comparable with the injection scale. We provide a physical explanation of the existence of such a two-dimensional to three-dimensional correspondence and discuss the uncertainties in evaluating the damping scale of ions that can be obtained from <span class="hlt">observations</span>. However, we show that the statistics of <span class="hlt">velocity</span> dispersion from <span class="hlt">observed</span> line profiles can provide the spectral index and the energy transfer rate of turbulence. Also</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19810007123','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810007123"><span>A proposed method for wind <span class="hlt">velocity</span> measurement from space</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Censor, D.; Levine, D. M.</p> <p>1980-01-01</p> <p>An investigation was made of the feasibility of making wind <span class="hlt">velocity</span> measurements from space by monitoring the apparent change in the refractive index of the atmosphere induced by motion of the air. The physical principle is the same as that resulting in the phase changes measured in the Fizeau experiment. It is proposed that this phase change could be measured using a three cornered arrangement of satellite borne source and reflectors, around which two laser beams propagate in opposite directions. It is shown that even though the <span class="hlt">velocity</span> of the satellites is much larger than the wind <span class="hlt">velocity</span>, factors such as change in satellite position and Doppler shifts can be taken into account in a reasonable manner and the Fizeau phase measured. This phase measurement yields an average wind <span class="hlt">velocity</span> along the ray path through the atmosphere. The method requires neither high accuracy for satellite position or <span class="hlt">velocity</span>, nor precise knowledge of the refractive index or its <span class="hlt">gradient</span> in the atmosphere. However, the method intrinsically yields wind <span class="hlt">velocity</span> integrated along the ray path; hence to obtain higher spatial resolution, inversion techniques are required.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28878264','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28878264"><span>Manipulation of acoustic wavefront by <span class="hlt">gradient</span> metasurface based on Helmholtz Resonators.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lan, Jun; Li, Yifeng; Xu, Yue; Liu, Xiaozhou</p> <p>2017-09-06</p> <p>We designed a <span class="hlt">gradient</span> acoustic metasurface to manipulate acoustic wavefront freely. The broad bandwidth and high efficiency transmission are achieved by the acoustic metasurface which is constructed with a series of unit cells to provide desired discrete acoustic <span class="hlt">velocity</span> distribution. Each unit cell is composed of a decorated metal plate with four periodically arrayed Helmholtz resonators (HRs) and a single slit. The design employs a <span class="hlt">gradient</span> <span class="hlt">velocity</span> to redirect refracted wave and the impedance matching between the metasurface and the background medium can be realized by adjusting the slit width of unit cell. The theoretical and numerical results show that some excellent wavefront manipulations are demonstrated by anomalous refraction, non-diffracting Bessel beam, sub-wavelength flat focusing, and effective tunable acoustic negative refraction. Our designed structure may offer potential applications for the imaging system, beam steering and acoustic lens.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29303327','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29303327"><span>Canceling the Gravity <span class="hlt">Gradient</span> Phase Shift in Atom Interferometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>D'Amico, G; Rosi, G; Zhan, S; Cacciapuoti, L; Fattori, M; Tino, G M</p> <p>2017-12-22</p> <p>Gravity <span class="hlt">gradients</span> represent a major obstacle in high-precision measurements by atom interferometry. Controlling their effects to the required stability and accuracy imposes very stringent requirements on the relative positioning of freely falling atomic clouds, as in the case of precise tests of Einstein's equivalence principle. We demonstrate a new method to exactly compensate the effects introduced by gravity <span class="hlt">gradients</span> in a Raman-pulse atom interferometer. By shifting the frequency of the Raman lasers during the central π pulse, it is possible to cancel the initial position- and <span class="hlt">velocity</span>-dependent phase shift produced by gravity <span class="hlt">gradients</span>. We apply this technique to simultaneous interferometers positioned along the vertical direction and demonstrate a new method for measuring local gravity <span class="hlt">gradients</span> that does not require precise knowledge of the relative position between the atomic clouds. Based on this method, we also propose an improved scheme to determine the Newtonian gravitational constant G towards the 10 ppm relative uncertainty.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvL.119y3201D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvL.119y3201D"><span>Canceling the Gravity <span class="hlt">Gradient</span> Phase Shift in Atom Interferometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>D'Amico, G.; Rosi, G.; Zhan, S.; Cacciapuoti, L.; Fattori, M.; Tino, G. M.</p> <p>2017-12-01</p> <p>Gravity <span class="hlt">gradients</span> represent a major obstacle in high-precision measurements by atom interferometry. Controlling their effects to the required stability and accuracy imposes very stringent requirements on the relative positioning of freely falling atomic clouds, as in the case of precise tests of Einstein's equivalence principle. We demonstrate a new method to exactly compensate the effects introduced by gravity <span class="hlt">gradients</span> in a Raman-pulse atom interferometer. By shifting the frequency of the Raman lasers during the central π pulse, it is possible to cancel the initial position- and <span class="hlt">velocity</span>-dependent phase shift produced by gravity <span class="hlt">gradients</span>. We apply this technique to simultaneous interferometers positioned along the vertical direction and demonstrate a new method for measuring local gravity <span class="hlt">gradients</span> that does not require precise knowledge of the relative position between the atomic clouds. Based on this method, we also propose an improved scheme to determine the Newtonian gravitational constant G towards the 10 ppm relative uncertainty.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3220692','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3220692"><span>Do Patterns of Bacterial Diversity along Salinity <span class="hlt">Gradients</span> Differ from Those <span class="hlt">Observed</span> for Macroorganisms?</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhang, Yong; Shen, Ji; van der Gast, Christopher; Hahn, Martin W.; Wu, Qinglong</p> <p>2011-01-01</p> <p>It is widely accepted that biodiversity is lower in more extreme environments. In this study, we sought to determine whether this trend, well documented for macroorganisms, also holds at the microbial level for bacteria. We used denaturing <span class="hlt">gradient</span> gel electrophoresis (DGGE) with phylum-specific primers to quantify the taxon richness (i.e., the DGGE band numbers) of the bacterioplankton communities of 32 pristine Tibetan lakes that represent a broad salinity range (freshwater to hypersaline). For the lakes investigated, salinity was found to be the environmental variable with the strongest influence on the bacterial community composition. We found that the bacterial taxon richness in freshwater habitats increased with increasing salinity up to a value of 1‰. In saline systems (systems with >1‰ salinity), the expected decrease of taxon richness along a <span class="hlt">gradient</span> of further increasing salinity was not <span class="hlt">observed</span>. These patterns were consistently <span class="hlt">observed</span> for two sets of samples taken in two different years. A comparison of 16S rRNA gene clone libraries revealed that the bacterial community of the lake with the highest salinity was characterized by a higher recent accelerated diversification than the community of a freshwater lake, whereas the phylogenetic diversity in the hypersaline lake was lower than that in the freshwater lake. These results suggest that different evolutionary forces may act on bacterial populations in freshwater and hypersaline lakes on the Tibetan Plateau, potentially resulting in different community structures and diversity patterns. PMID:22125616</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004MNRAS.351..265D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004MNRAS.351..265D"><span>The cluster galaxy circular <span class="hlt">velocity</span> function</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Desai, V.; Dalcanton, J. J.; Mayer, L.; Reed, D.; Quinn, T.; Governato, F.</p> <p>2004-06-01</p> <p>We present galaxy circular <span class="hlt">velocity</span> functions (GCVFs) for 34 low-redshift (z<~ 0.15) clusters identified in the Sloan Digital Sky Survey (SDSS), for 15 clusters drawn from dark matter simulations of hierarchical structure growth in a ΛCDM cosmology, and for ~22 000 SDSS field galaxies. We find that the simulations successfully reproduce the shape, amplitude and scatter in the <span class="hlt">observed</span> distribution of cluster galaxy circular <span class="hlt">velocities</span>. The power-law slope of the <span class="hlt">observed</span> cluster GCVF is ~-2.4, independent of cluster <span class="hlt">velocity</span> dispersion. The average slope of the simulated GCVFs is somewhat steeper, although formally consistent given the errors. We find that the effects of baryons on galaxy rotation curves is to flatten the simulated cluster GCVF into better agreement with <span class="hlt">observations</span>. The cumulative GCVFs of the simulated clusters are very similar across a wide range of cluster masses, provided individual subhalo circular <span class="hlt">velocities</span> are scaled by the circular <span class="hlt">velocities</span> of the parent cluster. The scatter is consistent with that measured in the cumulative, scaled <span class="hlt">observed</span> cluster GCVF. Finally, the <span class="hlt">observed</span> field GCVF deviates significantly from a power law, being flatter than the cluster GCVF at circular <span class="hlt">velocities</span> less than 200 km s-1.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RSOS....572421T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RSOS....572421T"><span>Collective cell migration without proliferation: density determines cell <span class="hlt">velocity</span> and wave <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tlili, Sham; Gauquelin, Estelle; Li, Brigitte; Cardoso, Olivier; Ladoux, Benoît; Delanoë-Ayari, Hélène; Graner, François</p> <p>2018-05-01</p> <p>Collective cell migration contributes to embryogenesis, wound healing and tumour metastasis. Cell monolayer migration experiments help in understanding what determines the movement of cells far from the leading edge. Inhibiting cell proliferation limits cell density increase and prevents jamming; we <span class="hlt">observe</span> long-duration migration and quantify space-time characteristics of the <span class="hlt">velocity</span> profile over large length scales and time scales. <span class="hlt">Velocity</span> waves propagate backwards and their frequency depends only on cell density at the moving front. Both cell average <span class="hlt">velocity</span> and wave <span class="hlt">velocity</span> increase linearly with the cell effective radius regardless of the distance to the front. Inhibiting lamellipodia decreases cell <span class="hlt">velocity</span> while waves either disappear or have a lower frequency. Our model combines conservation laws, monolayer mechanical properties and a phenomenological coupling between strain and polarity: advancing cells pull on their followers, which then become polarized. With reasonable values of parameters, this model agrees with several of our experimental <span class="hlt">observations</span>. Together, our experiments and model disantangle the respective contributions of active <span class="hlt">velocity</span> and of proliferation in monolayer migration, explain how cells maintain their polarity far from the moving front, and highlight the importance of strain-polarity coupling and density in long-range information propagation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPCM...29E5301D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPCM...29E5301D"><span>Localization of massless Dirac particles via spatial modulations of the Fermi <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Downing, C. A.; Portnoi, M. E.</p> <p>2017-08-01</p> <p>The electrons found in Dirac materials are notorious for being difficult to manipulate due to the Klein phenomenon and absence of backscattering. Here we investigate how spatial modulations of the Fermi <span class="hlt">velocity</span> in two-dimensional Dirac materials can give rise to localization effects, with either full (zero-dimensional) confinement or partial (one-dimensional) confinement possible depending on the geometry of the <span class="hlt">velocity</span> modulation. We present several exactly solvable models illustrating the nature of the bound states which arise, revealing how the <span class="hlt">gradient</span> of the Fermi <span class="hlt">velocity</span> is crucial for determining fundamental properties of the bound states such as the zero-point energy. We discuss the implications for guiding electronic waves in few-mode waveguides formed by Fermi <span class="hlt">velocity</span> modulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720010338','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720010338"><span>FORTRAN program for calculating <span class="hlt">velocities</span> in the meridional plane of a turbomachine 1: Centrifugal compressor</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vanco, M. R.</p> <p>1972-01-01</p> <p>The program will determine the <span class="hlt">velocities</span> in the meridional plane of a backward-swept impeller, a radial impeller, and a vaned diffuser. The <span class="hlt">velocity</span> <span class="hlt">gradient</span> equation with the assumption of a hub-to-shroud mean stream surface is solved along arbitrary quasi-orthogonals in the meridional plane. These quasi-orthogonals are fixed straight lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFMDI14A..08H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFMDI14A..08H"><span><span class="hlt">Velocities</span> of Subducted Sediments and Continents</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hacker, B. R.; van Keken, P. E.; Abers, G. A.; Seward, G.</p> <p>2009-12-01</p> <p>The growing capability to measure seismic <span class="hlt">velocities</span> in subduction zones has led to unusual <span class="hlt">observations</span>. For example, although most minerals have VP/ VS ratios around 1.77, ratios <1.7 and >1.8 have been <span class="hlt">observed</span>. Here we explore the <span class="hlt">velocities</span> of subducted sediments and continental crust from trench to sub-arc depths using two methods. (1) Mineralogy was calculated as a function of P & T for a range of subducted sediment compositions using Perple_X, and rock <span class="hlt">velocities</span> were calculated using the methodology of Hacker & Abers [2004]. Calculated slab-top temperatures have 3 distinct depth intervals with different dP/dT <span class="hlt">gradients</span> that are determined by how coupling between the slab and mantle wedge is modeled. These three depth intervals show concomitant changes in VP and VS: <span class="hlt">velocities</span> initially increase with depth, then decrease beyond the modeled decoupling depth where induced flow in the wedge causes rapid heating, and increase again at depth. Subducted limestones, composed chiefly of aragonite, show monotonic increases in VP/ VS from 1.63 to 1.72. Cherts show large jumps in VP/ VS from 1.55-1.65 to 1.75 associated with the quartz-coesite transition. Terrigenous sediments dominated by quartz and mica show similar, but more-subdued, transitions from ~1.67 to 1.78. Pelagic sediments dominated by mica and clinopyroxene show near-monotonic increases in VP/ VS from 1.74 to 1.80. Subducted continental crust that is too dry to transform to high-pressure minerals has a VP/ VS ratio of 1.68-1.70. (2) <span class="hlt">Velocity</span> anisotropy calculations were made for the same P-T dependent mineralogies using the Christoffel equation and crystal preferred orientations measured via electron-backscatter diffraction for typical constituent phases. The calculated <span class="hlt">velocity</span> anisotropies range from 5-30%. For quartz-rich rocks, the calculated <span class="hlt">velocities</span> show a distinct depth dependence because crystal slip systems and CPOs change with temperature. In such rocks, the fast VP direction varies from slab</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1983JGR....88.9341B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1983JGR....88.9341B"><span>Sediment sound <span class="hlt">velocities</span> from Sonobuoys: Sunda Trench and forearc basins, Nicobar and Central Bengal Fans, and Andaman Sea Basins</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bachman, Richard T.; Hamilton, Edwin L.; Curray, Joseph R.</p> <p>1983-11-01</p> <p>Supplement is available with entire article on microfiche. Order from American Geophysical Union, 2000 Florida Avenue, N.W., Washington, DC 20009. Document B83-007; $2.50. Payment must accompany order. Measurements of mean sound <span class="hlt">velocities</span> in the first, largely unlithified layers in the seafloor were made using the sonobuoy technique in several areas in the northern Indian Ocean. Older measurements were added to new measurements, and regressions for mean and instantaneous <span class="hlt">velocity</span> versus one-way travel time of sound are presented for the central Bengal Fan, the central Andaman Sea Basin, the Nicobar Fan, and the Sunda Trench. New data and regression equations are presented for the Mergui-north Sumatra Basin and for four forearc basins between Sumatra and Java and the Sunda Trench. Minimum <span class="hlt">velocity</span> <span class="hlt">gradients</span> were found in those areas where sedimentation rates were high, and sediments have accumulated in thick sections which have not had time to fully consolidate (porosity in the top of the sediment section has not been fully reduced under overburden pressure). These minimum <span class="hlt">velocity</span> <span class="hlt">gradients</span> (just under the seafloor) were found in the four forearc basins where they ranged from 0.34 s-1 to 0.84 s-1 with an average of 0.58 s-1. The near-surface <span class="hlt">velocity</span> <span class="hlt">gradient</span> in the Sunda Trench was 1.33 s-1, but was higher in the adjacent, fossil Nicobar Fan (1.62 s-1). In the surface of the Bengal Fan the <span class="hlt">velocity</span> <span class="hlt">gradient</span> was low in the upper fan (0.86 s-1), high in the central fan (1.94 s-1), and again lower in the southern fan (1.18 s-1), which may support sedimentation models calling for bypassing of the central fan and higher rates of accumulation on the southern fan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890022645&hterms=Dwarf+stars+blue&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DWhy%2BDwarf%2Bstars%2Bblue','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890022645&hterms=Dwarf+stars+blue&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DWhy%2BDwarf%2Bstars%2Bblue"><span>H-alpha Fabry-Perot interferometric <span class="hlt">observations</span> of blue compact dwarf galaxies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Thuan, Trinh Xuan; Williams, T. B.; Malumuth, E.</p> <p>1987-01-01</p> <p>H-alpha Fabry-Perot interferometric <span class="hlt">observations</span> of the two blue compact dwarf galaxies (BCDs) 7 Zw 403 and 1 Zw 49 are presented. The <span class="hlt">velocity</span> field of 7 Zw 403 shows no clear large-scale organized motion but the <span class="hlt">velocity</span> field is not completely chaotic either. The gas associated with the 8 H II regions in 7 Zw 403 has neither the highest nor lowest <span class="hlt">velocities</span>. The BCD 1 Zw 49 is dominated by a single H II region which is about 50 times brighter than any other feature in the galaxy. There is a chain of fainter H II regions extending across the galaxy. The <span class="hlt">velocity</span> field is well ordered along the H II region chain, but it is very complex around the dominant H II region, suggesting H-alpha loops and filaments around the latter. Both BCDs show <span class="hlt">velocity</span> <span class="hlt">gradients</span> of about 25 km/s on scales of about 10 pc in 7 Zw 403 and of about 50 pc in 1 Zw 49. These <span class="hlt">velocity</span> discontinuities compress the gas and are probably responsible for the star formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16958074','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16958074"><span>Quantification of intravoxel <span class="hlt">velocity</span> standard deviation and turbulence intensity by generalizing phase-contrast MRI.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dyverfeldt, Petter; Sigfridsson, Andreas; Kvitting, John-Peder Escobar; Ebbers, Tino</p> <p>2006-10-01</p> <p>Turbulent flow, characterized by <span class="hlt">velocity</span> fluctuations, is a contributing factor to the pathogenesis of several cardiovascular diseases. A clinical noninvasive tool for assessing turbulence is lacking, however. It is well known that the occurrence of multiple spin <span class="hlt">velocities</span> within a voxel during the influence of a magnetic <span class="hlt">gradient</span> moment causes signal loss in phase-contrast magnetic resonance imaging (PC-MRI). In this paper a mathematical derivation of an expression for computing the standard deviation (SD) of the blood flow <span class="hlt">velocity</span> distribution within a voxel is presented. The SD is obtained from the magnitude of PC-MRI signals acquired with different first <span class="hlt">gradient</span> moments. By exploiting the relation between the SD and turbulence intensity (TI), this method allows for quantitative studies of turbulence. For validation, the TI in an in vitro flow phantom was quantified, and the results compared favorably with previously published laser Doppler anemometry (LDA) results. This method has the potential to become an important tool for the noninvasive assessment of turbulence in the arterial tree.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22126970-metallicity-distribution-functions-radial-velocities-alpha-element-abundances-three-off-axis-bulge-fields','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22126970-metallicity-distribution-functions-radial-velocities-alpha-element-abundances-three-off-axis-bulge-fields"><span>METALLICITY DISTRIBUTION FUNCTIONS, RADIAL <span class="hlt">VELOCITIES</span>, AND ALPHA ELEMENT ABUNDANCES IN THREE OFF-AXIS BULGE FIELDS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Johnson, Christian I.; Rich, R. Michael; Kobayashi, Chiaki</p> <p>2013-03-10</p> <p>We present radial <span class="hlt">velocities</span> and chemical abundance ratios of [Fe/H], [O/Fe], [Si/Fe], and [Ca/Fe] for 264 red giant branch stars in three Galactic bulge off-axis fields located near (l, b) = (-5.5, -7), (-4, -9), and (+8.5, +9). The results are based on equivalent width and spectrum synthesis analyses of moderate resolution (R Almost-Equal-To 18,000), high signal-to-noise ratio (S/N {approx} 75-300 pixel{sup -1}) spectra obtained with the Hydra spectrographs on the Blanco 4 m and WIYN 3.5 m telescopes. The targets were selected from the blue side of the giant branch to avoid cool stars that would be strongly affectedmore » by CN and TiO; however, a comparison of the color-metallicity distribution in literature samples suggests that our selection of bluer targets should not present a significant bias against metal-rich stars. We find a full range in metallicity that spans [Fe/H] Almost-Equal-To -1.5 to +0.5, and that, in accordance with the previously <span class="hlt">observed</span> minor-axis vertical metallicity <span class="hlt">gradient</span>, the median [Fe/H] also declines with increasing Galactic latitude in off-axis fields. The off-axis vertical [Fe/H] <span class="hlt">gradient</span> in the southern bulge is estimated to be {approx}0.4 dex kpc{sup -1}; however, comparison with the minor-axis data suggests that a strong radial <span class="hlt">gradient</span> does not exist. The (+8.5, +9) field exhibits a higher than expected metallicity, with a median [Fe/H] = -0.23, that might be related to a stronger presence of the X-shaped bulge structure along that line-of-sight. This could also be the cause of an anomalous increase in the median radial <span class="hlt">velocity</span> for intermediate metallicity stars in the (+8.5, +9) field. However, the overall radial <span class="hlt">velocity</span> and dispersion for each field are in good agreement with recent surveys and bulge models. All fields exhibit an identical, strong decrease in <span class="hlt">velocity</span> dispersion with increasing metallicity that is consistent with <span class="hlt">observations</span> in similar minor-axis outer bulge fields. Additionally, the [O/Fe], [Si</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA12A..06Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA12A..06Y"><span>Vertical Rise <span class="hlt">Velocity</span> of Equatorial Plasma Bubbles Estimated from Equatorial Atmosphere Radar <span class="hlt">Observations</span> and High-Resolution Bubble Model Simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yokoyama, T.; Ajith, K. K.; Yamamoto, M.; Niranjan, K.</p> <p>2017-12-01</p> <p>Equatorial plasma bubble (EPB) is a well-known phenomenon in the equatorial ionospheric F region. As it causes severe scintillation in the amplitude and phase of radio signals, it is important to understand and forecast the occurrence of EPBs from a space weather point of view. The development of EPBs is presently believed as an evolution of the generalized Rayleigh-Taylor instability. We have already developed a 3D high-resolution bubble (HIRB) model with a grid spacing of as small as 1 km and presented nonlinear growth of EPBs which shows very turbulent internal structures such as bifurcation and pinching. As EPBs have field-aligned structures, the latitude range that is affected by EPBs depends on the apex altitude of EPBs over the dip equator. However, it was not easy to <span class="hlt">observe</span> the apex altitude and vertical rise <span class="hlt">velocity</span> of EPBs. Equatorial Atmosphere Radar (EAR) in Indonesia is capable of steering radar beams quickly so that the growth phase of EPBs can be captured clearly. The vertical rise <span class="hlt">velocities</span> of the EPBs <span class="hlt">observed</span> around the midnight hours are significantly smaller compared to those <span class="hlt">observed</span> in postsunset hours. Further, the vertical growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The HIRB model with varying background conditions are employed to investigate the possible factors that control the vertical rise <span class="hlt">velocity</span> and maximum attainable altitudes of EPBs. The estimated rise <span class="hlt">velocities</span> from EAR <span class="hlt">observations</span> at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20033047','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20033047"><span>The <span class="hlt">velocity</span> of climate change.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Loarie, Scott R; Duffy, Philip B; Hamilton, Healy; Asner, Gregory P; Field, Christopher B; Ackerly, David D</p> <p>2009-12-24</p> <p>The ranges of plants and animals are moving in response to recent changes in climate. As temperatures rise, ecosystems with 'nowhere to go', such as mountains, are considered to be more threatened. However, species survival may depend as much on keeping pace with moving climates as the climate's ultimate persistence. Here we present a new index of the <span class="hlt">velocity</span> of temperature change (km yr(-1)), derived from spatial <span class="hlt">gradients</span> ( degrees C km(-1)) and multimodel ensemble forecasts of rates of temperature increase ( degrees C yr(-1)) in the twenty-first century. This index represents the instantaneous local <span class="hlt">velocity</span> along Earth's surface needed to maintain constant temperatures, and has a global mean of 0.42 km yr(-1) (A1B emission scenario). Owing to topographic effects, the <span class="hlt">velocity</span> of temperature change is lowest in mountainous biomes such as tropical and subtropical coniferous forests (0.08 km yr(-1)), temperate coniferous forest, and montane grasslands. <span class="hlt">Velocities</span> are highest in flooded grasslands (1.26 km yr(-1)), mangroves and deserts. High <span class="hlt">velocities</span> suggest that the climates of only 8% of global protected areas have residence times exceeding 100 years. Small protected areas exacerbate the problem in Mediterranean-type and temperate coniferous forest biomes. Large protected areas may mitigate the problem in desert biomes. These results indicate management strategies for minimizing biodiversity loss from climate change. Montane landscapes may effectively shelter many species into the next century. Elsewhere, reduced emissions, a much expanded network of protected areas, or efforts to increase species movement may be necessary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21392527-effects-non-circular-motions-azimuthal-color-gradients','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21392527-effects-non-circular-motions-azimuthal-color-gradients"><span>EFFECTS OF NON-CIRCULAR MOTIONS ON AZIMUTHAL COLOR <span class="hlt">GRADIENTS</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Martinez-Garcia, Eric E.; Gonzalez-Lopezlira, Rosa A.; Gomez, Gilberto C., E-mail: emartinez@cida.v, E-mail: r.gonzalez@crya.unam.m, E-mail: g.gomez@crya.unam.m</p> <p>2009-12-20</p> <p>Assuming that density waves trigger star formation, and that young stars preserve the <span class="hlt">velocity</span> components of the molecular gas where they are born, we analyze the effects that non-circular gas orbits have on color <span class="hlt">gradients</span> across spiral arms. We try two approaches, one involving semianalytical solutions for spiral shocks, and another with magnetohydrodynamic (MHD) numerical simulation data. We find that, if non-circular motions are ignored, the comparison between <span class="hlt">observed</span> color <span class="hlt">gradients</span> and stellar population synthesis models would in principle yield pattern speed values that are systematically too high for regions inside corotation, with the difference between the real and themore » measured pattern speeds increasing with decreasing radius. On the other hand, image processing and pixel averaging result in systematically lower measured spiral pattern speed values, regardless of the kinematics of stellar orbits. The net effect is that roughly the correct pattern speeds are recovered, although the trend of higher measured OMEGA{sub p} at lower radii (as expected when non-circular motions exist but are neglected) should still be <span class="hlt">observed</span>. We examine the MartInez-GarcIa et al. photometric data and confirm that this is indeed the case. The comparison of the size of the systematic pattern speed offset in the data with the predictions of the semianalytical and MHD models corroborates that spirals are more likely to end at outer Lindblad resonance, as these authors had already found.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998JGR...10324321L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998JGR...10324321L"><span>Quasi-Love phases between Tonga and Hawaii: <span class="hlt">Observations</span>, simulations, and explanations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Levin, Vadim; Park, Jeffrey</p> <p>1998-10-01</p> <p>Seismograms of some shallow Tonga earthquakes <span class="hlt">observed</span> at Hawaii contain SV-polarized phases in the Love wave time window, most prominently on the vertical component. Given the geometry of the <span class="hlt">observations</span> (Δ ≈ 40-45°), such phases may be explained either as body waves or as mode-converted surface waves. Detailed synthetic seismogram modeling of representative events reveals several instances where the body wave explanation is inadequate, even when plausible uncertainties in the source mechanism are taken into account. The <span class="hlt">observed</span>, SV-polarized phase can instead be generated through Love-Rayleigh scattering, which requires laterally varying seismic anisotropy along the Tonga-Hawaii path. Trial-and-error forward modeling with simple structures based on the transversely isotropic mid-Pacific <span class="hlt">velocity</span> model PA5 of Gaherty et al [1996] obtains <span class="hlt">velocity</span> structure that yields synthetic seismograms matching the <span class="hlt">observations</span>. This model, while non unique, suggests first-order constraints on the lateral variation in anisotropic properties, and associated mantle flow, along the Tonga-Hawaii path. By examining trade-offs in model parameters, we conclude that robust features of the model are: (1) a transition from radial to mixed radial and azimuthal anisotropy 3°-5° from Hawaii; (2) the NW-SE alignment of the axis of azimuthal anisotropy; (3) higher degree of P anisotropy relative to S anisotropy; and (4) the presence of azimuthal anisotropy within upper 200-250 km of the mantle. Taken together, these features imply a disruption of mantle fabric by the processes forming Hawaii-Emperor volcanic system. A model with anisotropic <span class="hlt">gradients</span> in both the lithospheric lid and shallow asthenosphere is the simplest extension of our starting model. However, an equivalent data fit can be obtained if the azimuthal-anisotropy <span class="hlt">gradients</span> are restricted to line beneath the high-<span class="hlt">velocity</span> "lid" of model PA5, so that mantle hot spot flow need not penetrate the lithospheric lid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMDI11B2587P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMDI11B2587P"><span>Shear wave <span class="hlt">velocity</span> structure in the lithosphere and asthenosphere across the Southern California continent and Pacific plate margin using inversion of Rayleigh wave data from the ALBACORE project.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Price, A. C.; Weeraratne, D. S.; Kohler, M. D.; Rathnayaka, S.; Escobar, L., Sr.</p> <p>2015-12-01</p> <p>The North American and Pacific plate boundary is a unique example of past subduction of an oceanic spreading center which has involved oceanic plate capture and inception of a continental transform boundary that juxtaposes continental and oceanic lithosphere on a single plate. The amphibious ALBACORE seismic project (Asthenospheric and Lithospheric Broadband Architecture from the California Offshore Region Experiment) deployed 34 ocean bottom seismometers (OBS) on 15-35 Ma seafloor and offers a unique opportunity to compare the LAB in continental and oceanic lithosphere in one seismic study. Rayleigh waves were recorded simultaneously by our offshore array and 82 CISN network land stations from 2010-2011. Here we predict phase <span class="hlt">velocities</span> for a starting shear wave <span class="hlt">velocity</span> model for each of 5 regions in our study area and compare to <span class="hlt">observed</span> phase <span class="hlt">velocities</span> from our array in a least-squares sense that produces the best fit 1-D shear wave <span class="hlt">velocity</span> structure for each region. Preliminary results for the deep ocean (seafloor 25-32 Ma) indicates high <span class="hlt">velocities</span> reaching 4.5 km/s at depths of 50 km associated with the lithosphere for seafloor 25-32 Ma. A negative <span class="hlt">velocity</span> <span class="hlt">gradient</span> is <span class="hlt">observed</span> below this which reaches a minimum of 4.0 km/s at 160 km depth. The mid-ocean region (age 13-25 Ma) indicates a slightly lower magnitude and shallower LVZ. The Inner Borderland displays the highest lithospheric <span class="hlt">velocities</span> offshore reaching 4.8 km/s at 40 km depth indicating underplating. The base of the LVZ in the Borderland increases sharply from 4.0 km/s to 4.5 km/s at 80-150 km depth indicating partial melt and compositional changes. The LVZ displays a very gradual positive <span class="hlt">velocity</span> <span class="hlt">gradient</span> in all other regions such as the deep seafloor and continent reaching 4.5 km/s at 300 km depth. The deep ocean, Borderlands, and continental region each have unique lithospheric <span class="hlt">velocities</span>, LAB depths, and LVZ character that indicate stark differences in mantle structure that occur on a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27031956','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27031956"><span>Dynamics of Scroll Wave in a Three-Dimensional System with Changing <span class="hlt">Gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yuan, Xiao-Ping; Chen, Jiang-Xing; Zhao, Ye-Hua; Liu, Gui-Quan; Ying, He-Ping</p> <p>2016-01-01</p> <p>The dynamics of a scroll wave in an excitable medium with <span class="hlt">gradient</span> excitability is studied in detail. Three parameter regimes can be distinguished by the degree of <span class="hlt">gradient</span>. For a small <span class="hlt">gradient</span>, the system reaches a simple rotating synchronization. In this regime, the rigid rotating <span class="hlt">velocity</span> of spiral waves is maximal in the layers with the highest filament twist. As the excitability <span class="hlt">gradient</span> increases, the scroll wave evolutes into a meandering synchronous state. This transition is accompanied by a variation in twisting rate. Filament twisting may prevent the breakup of spiral waves in the bottom layers with a low excitability with which a spiral breaks in a 2D medium. When the <span class="hlt">gradient</span> is large enough, the twisted filament breaks up, which results in a semi-turbulent state where the lower part is turbulent while the upper part contains a scroll wave with a low twisting filament.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24115059','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24115059"><span>An algorithm to estimate unsteady and quasi-steady pressure fields from <span class="hlt">velocity</span> field measurements.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dabiri, John O; Bose, Sanjeeb; Gemmell, Brad J; Colin, Sean P; Costello, John H</p> <p>2014-02-01</p> <p>We describe and characterize a method for estimating the pressure field corresponding to <span class="hlt">velocity</span> field measurements such as those obtained by using particle image velocimetry. The pressure <span class="hlt">gradient</span> is estimated from a time series of <span class="hlt">velocity</span> fields for unsteady calculations or from a single <span class="hlt">velocity</span> field for quasi-steady calculations. The corresponding pressure field is determined based on median polling of several integration paths through the pressure <span class="hlt">gradient</span> field in order to reduce the effect of measurement errors that accumulate along individual integration paths. Integration paths are restricted to the nodes of the measured <span class="hlt">velocity</span> field, thereby eliminating the need for measurement interpolation during this step and significantly reducing the computational cost of the algorithm relative to previous approaches. The method is validated by using numerically simulated flow past a stationary, two-dimensional bluff body and a computational model of a three-dimensional, self-propelled anguilliform swimmer to study the effects of spatial and temporal resolution, domain size, signal-to-noise ratio and out-of-plane effects. Particle image velocimetry measurements of a freely swimming jellyfish medusa and a freely swimming lamprey are analyzed using the method to demonstrate the efficacy of the approach when applied to empirical data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeoRL..4310078S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeoRL..4310078S"><span>The shallow elastic structure of the lunar crust: New insights from seismic wavefield <span class="hlt">gradient</span> analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sollberger, David; Schmelzbach, Cedric; Robertsson, Johan O. A.; Greenhalgh, Stewart A.; Nakamura, Yosio; Khan, Amir</p> <p>2016-10-01</p> <p>Enigmatic lunar seismograms recorded during the Apollo 17 mission in 1972 have so far precluded the identification of shear-wave arrivals and hence the construction of a comprehensive elastic model of the shallow lunar subsurface. Here, for the first time, we extract shear-wave information from the Apollo active seismic data using a novel waveform analysis technique based on spatial seismic wavefield <span class="hlt">gradients</span>. The star-like recording geometry of the active seismic experiment lends itself surprisingly well to compute spatial wavefield <span class="hlt">gradients</span> and rotational ground motion as a function of time. These <span class="hlt">observables</span>, which are new to seismic exploration in general, allowed us to identify shear waves in the complex lunar seismograms, and to derive a new model of seismic compressional and shear-wave <span class="hlt">velocities</span> in the shallow lunar crust, critical to understand its lithology and constitution, and its impact on other geophysical investigations of the Moon's deep interior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018InvPr..34d5006Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018InvPr..34d5006Y"><span>Total variation regularization for seismic waveform inversion using an adaptive primal dual hybrid <span class="hlt">gradient</span> method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yong, Peng; Liao, Wenyuan; Huang, Jianping; Li, Zhenchuan</p> <p>2018-04-01</p> <p>Full waveform inversion is an effective tool for recovering the properties of the Earth from seismograms. However, it suffers from local minima caused mainly by the limited accuracy of the starting model and the lack of a low-frequency component in the seismic data. Because of the high <span class="hlt">velocity</span> contrast between salt and sediment, the relation between the waveform and <span class="hlt">velocity</span> perturbation is strongly nonlinear. Therefore, salt inversion can easily get trapped in the local minima. Since the <span class="hlt">velocity</span> of salt is nearly constant, we can make the most of this characteristic with total variation regularization to mitigate the local minima. In this paper, we develop an adaptive primal dual hybrid <span class="hlt">gradient</span> method to implement total variation regularization by projecting the solution onto a total variation norm constrained convex set, through which the total variation norm constraint is satisfied at every model iteration. The smooth background <span class="hlt">velocities</span> are first inverted and the perturbations are gradually obtained by successively relaxing the total variation norm constraints. Numerical experiment of the projection of the BP model onto the intersection of the total variation norm and box constraints has demonstrated the accuracy and efficiency of our adaptive primal dual hybrid <span class="hlt">gradient</span> method. A workflow is designed to recover complex salt structures in the BP 2004 model and the 2D SEG/EAGE salt model, starting from a linear <span class="hlt">gradient</span> model without using low-frequency data below 3 Hz. The salt inversion processes demonstrate that wavefield reconstruction inversion with a total variation norm and box constraints is able to overcome local minima and inverts the complex salt <span class="hlt">velocity</span> layer by layer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1177645-testing-thermal-gradient-driving-force-grain-boundary-migration-using-molecular-dynamics-simulations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1177645-testing-thermal-gradient-driving-force-grain-boundary-migration-using-molecular-dynamics-simulations"><span>Testing thermal <span class="hlt">gradient</span> driving force for grain boundary migration using molecular dynamics simulations</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bai, Xian-Ming; Zhang, Yongfeng; Tonks, Michael R.</p> <p>2015-02-01</p> <p>Strong thermal <span class="hlt">gradients</span> in low-thermal-conductivity ceramics may drive extended defects, such as grain boundaries and voids, to migrate in preferential directions. In this work, molecular dynamics simulations are conducted to study thermal <span class="hlt">gradient</span> driven grain boundary migration and to verify a previously proposed thermal <span class="hlt">gradient</span> driving force equation, using uranium dioxide as a model system. It is found that a thermal <span class="hlt">gradient</span> drives grain boundaries to migrate up the <span class="hlt">gradient</span> and the migration <span class="hlt">velocity</span> increases under a constant <span class="hlt">gradient</span> owing to the increase in mobility with temperature. Different grain boundaries migrate at very different rates due to their different intrinsicmore » mobilities. The extracted mobilities from the thermal <span class="hlt">gradient</span> driven simulations are compared with those calculated from two other well-established methods and good agreement between the three different methods is found, demonstrating that the theoretical equation of the thermal <span class="hlt">gradient</span> driving force is valid, although a correction of one input parameter should be made. The discrepancy in the grain boundary mobilities between modeling and experiments is also discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMOS14A..02F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMOS14A..02F"><span>Airborne microwave radar measurements of surface <span class="hlt">velocity</span> in a tidally-driven inlet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Farquharson, G.; Thomson, J. M.</p> <p>2012-12-01</p> <p>A miniaturized dual-beam along-track interferometric (ATI) synthetic aperture radar (SAR), capable of measuring two components of surface <span class="hlt">velocity</span> at high resolution, was operated during the 2012 Rivers and Inlets Experiment (RIVET) at the New River Inlet in North Carolina. The inlet is predominantly tidally-driven, with little upstream river discharge. Surface <span class="hlt">velocities</span> in the inlet and nearshore region were measured during ebb and flood tides during a variety of wind and offshore wave conditions. The radar-derived surface <span class="hlt">velocities</span> range from around ±2~m~s1 during times of maximum flow. We compare these radar-derived surface <span class="hlt">velocities</span> with surface <span class="hlt">velocities</span> measured with drifters. The accuracy of the radar-derived <span class="hlt">velocities</span> is investigated, especially in areas of large <span class="hlt">velocity</span> <span class="hlt">gradients</span> where along-track interferometric SAR can show significant differences with surface <span class="hlt">velocity</span>. The goal of this research is to characterize errors in along-track interferometric SAR <span class="hlt">velocity</span> so that ATI SAR measurements can be coupled with data assimilative modeling with the goal of developing the capability to adequately constrain nearshore models using remote sensing measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1234198','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1234198"><span><span class="hlt">Observation</span> of 690 MV m -1 Electron Accelerating <span class="hlt">Gradient</span> with a Laser-Driven Dielectric Microstructure</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wootton, K. P.; Wu, Z.; Cowan, B. M.</p> <p></p> <p>Acceleration of electrons using laser-driven dielectric microstructures is a promising technology for the miniaturization of particle accelerators. In this work, experimental results are presented of relativistic electron acceleration with 690±100 MVm -1 <span class="hlt">gradient</span>. This is a record-high accelerating <span class="hlt">gradient</span> for a dielectric microstructure accelerator, nearly doubling the previous record <span class="hlt">gradient</span>. To reach higher acceleration <span class="hlt">gradients</span> the present experiment employs 90 fs duration laser pulses.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=241540','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=241540"><span>Purification of Giardia muris cysts by <span class="hlt">velocity</span> sedimentation.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sauch, J F</p> <p>1984-01-01</p> <p>Giardia muris cysts were separated from fecal contaminants in primary isolates by unit gravity <span class="hlt">velocity</span> sedimentation. Crude isolates obtained by centrifugation over 1.0 M sucrose were overlaid onto a Percoll density <span class="hlt">gradient</span>, 1.01 to 1.03 g/ml. G. muris cysts were well separated from faster-sedimenting fecal debris and slower-sedimenting Spironucleus muris and bacteria in 1.5 h. PMID:6486790</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...600A..70A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...600A..70A"><span>Red giants <span class="hlt">observed</span> by CoRoT and APOGEE: The evolution of the Milky Way's radial metallicity <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anders, F.; Chiappini, C.; Minchev, I.; Miglio, A.; Montalbán, J.; Mosser, B.; Rodrigues, T. S.; Santiago, B. X.; Baudin, F.; Beers, T. C.; da Costa, L. N.; García, R. A.; García-Hernández, D. A.; Holtzman, J.; Maia, M. A. G.; Majewski, S.; Mathur, S.; Noels-Grotsch, A.; Pan, K.; Schneider, D. P.; Schultheis, M.; Steinmetz, M.; Valentini, M.; Zamora, O.</p> <p>2017-04-01</p> <p>Using combined asteroseismic and spectroscopic <span class="hlt">observations</span> of 418 red-giant stars close to the Galactic disc plane (6 kpc < RGal ≲ 13 kpc, | ZGal| < 0.3 kpc), we measure the age dependence of the radial metallicity distribution in the Milky Way's thin disc over cosmic time. The slope of the radial iron <span class="hlt">gradient</span> of the young red-giant population (-0.058 ± 0.008 [stat.] ±0.003 [syst.] dex/kpc) is consistent with recent Cepheid measurements. For stellar populations with ages of 1-4 Gyr the <span class="hlt">gradient</span> is slightly steeper, at a value of -0.066 ± 0.007 ± 0.002 dex/kpc, and then flattens again to reach a value of -0.03 dex/kpc for stars with ages between 6 and 10 Gyr. Our results are in good agreement with a state-of-the-art chemo-dynamical Milky-Way model in which the evolution of the abundance <span class="hlt">gradient</span> and its scatter can be entirely explained by a non-varying negative metallicity <span class="hlt">gradient</span> in the interstellar medium, together with stellar radial heating and migration. We also offer an explanation for why intermediate-age open clusters in the solar neighbourhood can be more metal-rich, and why their radial metallicity <span class="hlt">gradient</span> seems to be much steeper than that of the youngest clusters. Already within 2 Gyr, radial mixing can bring metal-rich clusters from the innermost regions of the disc to Galactocentric radii of 5 to 8 kpc. We suggest that these outward-migrating clusters may be less prone to tidal disruption and therefore steepen the local intermediate-age cluster metallicity <span class="hlt">gradient</span>. Our scenario also explains why the strong steepening of the local iron <span class="hlt">gradient</span> with age is not seen in field stars. In the near future, asteroseismic data from the K2 mission will allow for improved statistics and a better coverage of the inner-disc regions, thereby providing tighter constraints on theevolution of the central parts of the Milky Way.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830054010&hterms=Active+vials&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DActive%2Bvials','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830054010&hterms=Active+vials&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DActive%2Bvials"><span>A surge <span class="hlt">observed</span> in H alpha and C IV</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schmieder, B.; Mein, P.; Vial, J. C.; Tandberg-Hanssen, E.</p> <p>1982-01-01</p> <p>Results are presented of simultaneous measurements of H-alpha (MSDP at Meudon) and C IV (UVSP onboard SMM) of Active Region 2701 made on October 2, 1980. Isodensity and <span class="hlt">velocity</span> maps were obtained for both lines and these maps were superimposed. Results show a good correlation between the H-alpha and C IV <span class="hlt">velocities</span> with a surge being <span class="hlt">observed</span> for 10 minutes. The base of the surge was determined to be located in a bright point in C IV and H-alpha, while the escaping matter followed the same channel ('absorbing' in H-alpha, 'emitting' in C IV). It was found that the <span class="hlt">velocity</span> along the surge was about 80 km/s in H-alpha and 100 km/s in C IV. In addition, a loop appeared in C IV during the surge. It is concluded that the vertical pressure <span class="hlt">gradient</span> was capable of driving the surge.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25131340','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25131340"><span>Auditory <span class="hlt">velocity</span> discrimination in the horizontal plane at very high <span class="hlt">velocities</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Frissen, Ilja; Féron, François-Xavier; Guastavino, Catherine</p> <p>2014-10-01</p> <p>We determined <span class="hlt">velocity</span> discrimination thresholds and Weber fractions for sounds revolving around the listener at very high <span class="hlt">velocities</span>. Sounds used were a broadband white noise and two harmonic sounds with fundamental frequencies of 330 Hz and 1760 Hz. Experiment 1 used <span class="hlt">velocities</span> ranging between 288°/s and 720°/s in an acoustically treated room and Experiment 2 used <span class="hlt">velocities</span> between 288°/s and 576°/s in a highly reverberant hall. A third experiment addressed potential confounds in the first two experiments. The results show that people can reliably discriminate <span class="hlt">velocity</span> at very high <span class="hlt">velocities</span> and that both thresholds and Weber fractions decrease as <span class="hlt">velocity</span> increases. These results violate Weber's law but are consistent with the empirical trend <span class="hlt">observed</span> in the literature. While thresholds for the noise and 330 Hz harmonic stimulus were similar, those for the 1760 Hz harmonic stimulus were substantially higher. There were no reliable differences in <span class="hlt">velocity</span> discrimination between the two acoustical environments, suggesting that auditory motion perception at high <span class="hlt">velocities</span> is robust against the effects of reverberation. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950037150&hterms=801&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D801','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950037150&hterms=801&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D801"><span>On the temperature and <span class="hlt">velocity</span> through the photosphere of a sunspot penumbra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Del Toro Iniesta, J. C.; Tarbell, T. D.; Cobo, B. Ruiz</p> <p>1994-01-01</p> <p>We investigate the structure in depth of a sunspot penumbra by means of the inversion code of the radiative transfer equation proposed by Ruiz Cobo & del Toro Iniesta (1992), applied to a set of filtergrams of a sunspot, scanning the Fe I line at 5576.1 A, with a sampling interval of 30 mA, from -120 to 120 mA from line center (data previously analyzed by Title et al. 1993). The temperature structure of this penumbra is obtained for each of the 801 pixels selected (0.32 sec x 0.32 sec). On the average, the temperatures seem to decrease as we move inward, but the differences are of the order of the rms values (approximately equal 100-200 K) at a given distance to sunspot center. The outer parts of the penumbra have also a bigger curvature in the T versus log tau(sub 5) relation than the inner parts. We realize, however, that these differences might be influenced by possible stray light effects. Compared to the quiet Sun, penumbral temperatures are cooler at deep layers and hotter at high layers. A mean penumbral model atmosphere is presented. The asymmetries <span class="hlt">observed</span> in the intensity profile (the line is magnetically insensitive) are deduced to be produced by strong <span class="hlt">gradients</span> of the line-of-sight <span class="hlt">velocity</span> that sharply vary spatially along slices of almost constant distance to sunspot center. These variations suggest that such <span class="hlt">gradients</span> are not only needed to explain the broadband circular polarization <span class="hlt">observed</span> in sunspots (see Sanchez Almeida & Lites 1992) but are a main characteristic of the fine-scale penumbra. The results are compatible with an Evershed flow present everywhere, but its <span class="hlt">gradient</span> with depth turns out to vary so that the flow seems to be mainly concentrated in some penumbral fibrils when studied through Dopplergrams. Finally, as by-products of this study, we put constraints to the practical usefulness of the Eddington-Barbier relation, and we explain the values of the Fourier Dopplergrams to be carrying information of layers around the centroid of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990092486&hterms=macmillan&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmacmillan','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990092486&hterms=macmillan&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmacmillan"><span>Global <span class="hlt">Velocities</span> from VLBI</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ma, Chopo; Gordon, David; MacMillan, Daniel</p> <p>1999-01-01</p> <p>Precise geodetic Very Long Baseline Interferometry (VLBI) measurements have been made since 1979 at about 130 points on all major tectonic plates, including stable interiors and deformation zones. From the data set of about 2900 <span class="hlt">observing</span> sessions and about 2.3 million <span class="hlt">observations</span>, useful three-dimensional <span class="hlt">velocities</span> can be derived for about 80 sites using an incremental least-squares adjustment of terrestrial, celestial, Earth rotation and site/session-specific parameters. The long history and high precision of the data yield formal errors for horizontal <span class="hlt">velocity</span> as low as 0.1 mm/yr, but the limitation on the interpretation of individual site <span class="hlt">velocities</span> is the tie to the terrestrial reference frame. Our studies indicate that the effect of converting precise relative VLBI <span class="hlt">velocities</span> to individual site <span class="hlt">velocities</span> is an error floor of about 0.4 mm/yr. Most VLBI horizontal <span class="hlt">velocities</span> in stable plate interiors agree with the NUVEL-1A model, but there are significant departures in Africa and the Pacific. Vertical precision is worse by a factor of 2-3, and there are significant non-zero values that can be interpreted as post-glacial rebound, regional effects, and local disturbances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DPPBI2003H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DPPBI2003H"><span>Direct evidence of stationary zonal flows and critical <span class="hlt">gradient</span> behavior for Er during formation of the edge pedestal in JET</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hillesheim, Jon</p> <p>2015-11-01</p> <p>High spatial resolution measurements with Doppler backscattering in JET have provided new insights into the development of the edge radial electric field during pedestal formation. The characteristics of Er have been studied as a function of density at 2.5 MA plasma current and 3 T toroidal magnetic field. We <span class="hlt">observe</span> fine-scale spatial structure in the edge Er well prior to the LH transition, consistent with stationary zonal flows. Zonal flows are a fundamental mechanism for the saturation of turbulence and this is the first direct evidence of stationary zonal flows in a tokamak. The radial wavelength of the zonal flows systematically decreases with density. The zonal flows are clearest in Ohmic conditions, weaker in L-mode, and absent in H-mode. Measurements also show that after neutral beam heating is applied, the edge Er builds up at a constant <span class="hlt">gradient</span> into the core during L-mode, at radii where Er is mainly due to toroidal <span class="hlt">velocity</span>. The local stability of <span class="hlt">velocity</span> shear driven turbulence, such as the parallel <span class="hlt">velocity</span> <span class="hlt">gradient</span> mode, will be assessed with gyrokinetic simulations. This critical Er shear persists across the LH transition into H-mode. Surprisingly, a reduction in the apparent magnitude of the Er well depth is <span class="hlt">observed</span> directly following the LH transition at high densities. Establishing the physics basis for the LH transition is important for projecting scalings to ITER and these <span class="hlt">observations</span> challenge existing models based on increased Er shear or strong zonal flows as the trigger for the transition. This work has been carried out within the framework of the EUROfusion Consortium and has received funding from the Euratom research and training programme 2014-2018 under grant agreement No 633053. The views and opinions expressed herein do not necessarily reflect those of the European Commission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23970568','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23970568"><span>Mass flow and <span class="hlt">velocity</span> profiles in Neurospora hyphae: partial plug flow dominates intra-hyphal transport.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Abadeh, Aryan; Lew, Roger R</p> <p>2013-11-01</p> <p>Movement of nuclei, mitochondria and vacuoles through hyphal trunks of Neurospora crassa were vector-mapped using fluorescent markers and green fluorescent protein tags. The vectorial movements of all three were strongly correlated, indicating the central role of mass (bulk) flow in cytoplasm movements in N. crassa. Profiles of <span class="hlt">velocity</span> versus distance from the hyphal wall did not match the parabolic shape predicted by the ideal Hagen-Poiseuille model of flow at low Reynolds number. Instead, the profiles were flat, consistent with a model of partial plug flow due to the high concentration of organelles in the flowing cytosol. The intra-hyphal pressure <span class="hlt">gradients</span> were manipulated by localized external osmotic treatments to demonstrate the dependence of <span class="hlt">velocity</span> (and direction) on pressure <span class="hlt">gradients</span> within the hyphae. The data support the concept that mass transport, driven by pressure <span class="hlt">gradients</span>, dominates intra-hyphal transport. The transport occurs by partial plug flow due to the organelles in the cytosol.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29251943','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29251943"><span>Controlling Directional Liquid Motion on Micro- and Nanocrystalline Diamond/β-SiC Composite <span class="hlt">Gradient</span> Films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Tao; Handschuh-Wang, Stephan; Huang, Lei; Zhang, Lei; Jiang, Xin; Kong, Tiantian; Zhang, Wenjun; Lee, Chun-Sing; Zhou, Xuechang; Tang, Yongbing</p> <p>2018-01-30</p> <p>In this Article, we report the synthesis of micro- and nanocrystalline diamond/β-SiC composite <span class="hlt">gradient</span> films, using a hot filament chemical vapor deposition (HFCVD) technique and its application as a robust and chemically inert means to actuate water and hazardous liquids. As revealed by scanning electron microscopy, the composition of the surface changed gradually from pure nanocrystalline diamond (hydrophobic) to a nanocrystalline β-SiC surface (hydrophilic). Transmission electron microscopy and Raman spectroscopy were employed to determine the presence of diamond, graphite, and β-SiC phases. The as-prepared <span class="hlt">gradient</span> films were evaluated for their ability to actuate water. Indeed, water was transported via the <span class="hlt">gradient</span> from the hydrophobic (hydrogen-terminated diamond) to the hydrophilic side (hydroxyl-terminated β-SiC) of the <span class="hlt">gradient</span> surface. The driving distance and <span class="hlt">velocity</span> of water is pivotally influenced by the surface roughness. The nanogradient surface showed significant promise as the lower roughness combined with the longer <span class="hlt">gradient</span> yields in transport distances of up to 3.7 mm, with a maximum droplet <span class="hlt">velocity</span> of nearly 250 mm/s measured by a high-speed camera. As diamond and β-SiC are chemically inert, the <span class="hlt">gradient</span> surfaces can be used to drive hazardous liquids and reactive mixtures, which was signified by the actuation of hydrochloric acid and sodium hydroxide solution. We envision that the diamond/β-SiC <span class="hlt">gradient</span> surface has high potential as an actuator for water transport in microfluidic devices, DNA sensors, and implants, which induce guided cell growth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004QJRMS.130.1977S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004QJRMS.130.1977S"><span>Agradient <span class="hlt">velocity</span>, vortical motion and gravity waves in a rotating shallow-water model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sutyrin Georgi, G.</p> <p>2004-07-01</p> <p>A new approach to modelling slow vortical motion and fast inertia-gravity waves is suggested within the rotating shallow-water primitive equations with arbitrary topography. The <span class="hlt">velocity</span> is exactly expressed as a sum of the <span class="hlt">gradient</span> wind, described by the Bernoulli function,B, and the remaining agradient part, proportional to the <span class="hlt">velocity</span> tendency. Then the equation for inverse potential vorticity,Q, as well as momentum equations for agradient <span class="hlt">velocity</span> include the same source of intrinsic flow evolution expressed as a single term J (B, Q), where J is the Jacobian operator (for any steady state J (B, Q) = 0). Two components of agradient <span class="hlt">velocity</span> are responsible for the fast inertia-gravity wave propagation similar to the traditionally used divergence and ageostrophic vorticity. This approach allows for the construction of balance relations for vortical dynamics and potential vorticity inversion schemes even for moderate Rossby and Froude numbers assuming the characteristic value of |J(B, Q)| = to be small. The components of agradient <span class="hlt">velocity</span> are used as the fast variables slaved to potential vorticity that allows for diagnostic estimates of the <span class="hlt">velocity</span> tendency, the direct potential vorticity inversion with the accuracy of 2 and the corresponding potential vorticity-conserving agradient <span class="hlt">velocity</span> balance model (AVBM). The ultimate limitations of constructing the balance are revealed in the form of the ellipticity condition for balanced tendency of the Bernoulli function which incorporates both known criteria of the formal stability: the <span class="hlt">gradient</span> wind modified by the characteristic vortical Rossby wave phase speed should be subcritical. The accuracy of the AVBM is illustrated by considering the linear normal modes and coastal Kelvin waves in the f-plane channel with topography.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApJ...744...14Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApJ...744...14Y"><span><span class="hlt">Velocity</span> Measurements for a Solar Active Region Fan Loop from Hinode/EIS <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Young, P. R.; O'Dwyer, B.; Mason, H. E.</p> <p>2012-01-01</p> <p>The <span class="hlt">velocity</span> pattern of a fan loop structure within a solar active region over the temperature range 0.15-1.5 MK is derived using data from the EUV Imaging Spectrometer (EIS) on board the Hinode satellite. The loop is aligned toward the <span class="hlt">observer</span>'s line of sight and shows downflows (redshifts) of around 15 km s-1 up to a temperature of 0.8 MK, but for temperatures of 1.0 MK and above the measured <span class="hlt">velocity</span> shifts are consistent with no net flow. This <span class="hlt">velocity</span> result applies over a projected spatial distance of 9 Mm and demonstrates that the cooler, redshifted plasma is physically disconnected from the hotter, stationary plasma. A scenario in which the fan loops consist of at least two groups of "strands"—one cooler and downflowing, the other hotter and stationary—is suggested. The cooler strands may represent a later evolutionary stage of the hotter strands. A density diagnostic of Mg VII was used to show that the electron density at around 0.8 MK falls from 3.2 × 109 cm-3 at the loop base, to 5.0 × 108 cm-3 at a projected height of 15 Mm. A filling factor of 0.2 is found at temperatures close to the formation temperature of Mg VII (0.8 MK), confirming that the cooler, downflowing plasma occupies only a fraction of the apparent loop volume. The fan loop is rooted within a so-called outflow region that displays low intensity and blueshifts of up to 25 km s-1 in Fe XII λ195.12 (formed at 1.5 MK), in contrast to the loop's redshifts of 15 km s-1 at 0.8 MK. A new technique for obtaining an absolute wavelength calibration for the EIS instrument is presented and an instrumental effect, possibly related to a distorted point-spread function, that affects <span class="hlt">velocity</span> measurements is identified.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22004363-velocity-measurements-solar-active-region-fan-loop-from-hinode-eis-observations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22004363-velocity-measurements-solar-active-region-fan-loop-from-hinode-eis-observations"><span><span class="hlt">VELOCITY</span> MEASUREMENTS FOR A SOLAR ACTIVE REGION FAN LOOP FROM HINODE/EIS <span class="hlt">OBSERVATIONS</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Young, P. R.; O'Dwyer, B.; Mason, H. E.</p> <p>2012-01-01</p> <p>The <span class="hlt">velocity</span> pattern of a fan loop structure within a solar active region over the temperature range 0.15-1.5 MK is derived using data from the EUV Imaging Spectrometer (EIS) on board the Hinode satellite. The loop is aligned toward the <span class="hlt">observer</span>'s line of sight and shows downflows (redshifts) of around 15 km s{sup -1} up to a temperature of 0.8 MK, but for temperatures of 1.0 MK and above the measured <span class="hlt">velocity</span> shifts are consistent with no net flow. This <span class="hlt">velocity</span> result applies over a projected spatial distance of 9 Mm and demonstrates that the cooler, redshifted plasma is physicallymore » disconnected from the hotter, stationary plasma. A scenario in which the fan loops consist of at least two groups of 'strands'-one cooler and downflowing, the other hotter and stationary-is suggested. The cooler strands may represent a later evolutionary stage of the hotter strands. A density diagnostic of Mg VII was used to show that the electron density at around 0.8 MK falls from 3.2 Multiplication-Sign 10{sup 9} cm{sup -3} at the loop base, to 5.0 Multiplication-Sign 10{sup 8} cm{sup -3} at a projected height of 15 Mm. A filling factor of 0.2 is found at temperatures close to the formation temperature of Mg VII (0.8 MK), confirming that the cooler, downflowing plasma occupies only a fraction of the apparent loop volume. The fan loop is rooted within a so-called outflow region that displays low intensity and blueshifts of up to 25 km s{sup -1} in Fe XII {lambda}195.12 (formed at 1.5 MK), in contrast to the loop's redshifts of 15 km s{sup -1} at 0.8 MK. A new technique for obtaining an absolute wavelength calibration for the EIS instrument is presented and an instrumental effect, possibly related to a distorted point-spread function, that affects <span class="hlt">velocity</span> measurements is identified.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhPl...24j3517S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhPl...24j3517S"><span>Analytic non-Maxwellian electron <span class="hlt">velocity</span> distribution function in a Hall discharge plasma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shagayda, Andrey; Tarasov, Alexey</p> <p>2017-10-01</p> <p>The electron <span class="hlt">velocity</span> distribution function in the low-pressure discharges with the crossed electric and magnetic fields, which occur in magnetrons, plasma accelerators, and Hall thrusters with a closed electron drift, is not Maxwellian. A deviation from equilibrium is caused by a large electron mean free path relative to the Larmor radius and the size of the discharge channel. In this study, we derived in the relaxation approximation the analytical expression of the electron <span class="hlt">velocity</span> distribution function in a weakly ionized Lorentz plasma with the crossed electric and magnetic fields in the presence of the electron density and temperature <span class="hlt">gradients</span> in the direction of the electric field. The solution was obtained in the stationary approximation far from boundary surfaces, when diffusion and mobility are determined by the classical effective collision frequency of electrons with ions and atoms. The moments of the distribution function including the average <span class="hlt">velocity</span>, the stress tensor, and the heat flux were calculated and compared with the classical hydrodynamic expressions. It was shown that a kinetic correction to the drift <span class="hlt">velocity</span> stems from a contribution of the off-diagonal component of the stress tensor. This correction becomes essential if the drift <span class="hlt">velocity</span> in the crossed electric and magnetic fields would be comparable to the thermal <span class="hlt">velocity</span> of electrons. The electron temperature has three different components at a nonzero effective collision frequency and two different components in the limit when the collision frequency tends to zero. It is shown that, in the presence of ionization collisions, the components of the heat flux have additives that are not related to the temperature <span class="hlt">gradient</span>, and arise because of the electron drift.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018A%26A...612A..84D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018A%26A...612A..84D"><span>Mapping the solar wind HI outflow <span class="hlt">velocity</span> in the inner heliosphere by coronagraphic ultraviolet and visible-light <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dolei, S.; Susino, R.; Sasso, C.; Bemporad, A.; Andretta, V.; Spadaro, D.; Ventura, R.; Antonucci, E.; Abbo, L.; Da Deppo, V.; Fineschi, S.; Focardi, M.; Frassetto, F.; Giordano, S.; Landini, F.; Naletto, G.; Nicolini, G.; Nicolosi, P.; Pancrazzi, M.; Romoli, M.; Telloni, D.</p> <p>2018-05-01</p> <p>We investigated the capability of mapping the solar wind outflow <span class="hlt">velocity</span> of neutral hydrogen atoms by using synergistic visible-light and ultraviolet <span class="hlt">observations</span>. We used polarised brightness images acquired by the LASCO/SOHO and Mk3/MLSO coronagraphs, and synoptic Lyα line <span class="hlt">observations</span> of the UVCS/SOHO spectrometer to obtain daily maps of solar wind H I outflow <span class="hlt">velocity</span> between 1.5 and 4.0 R⊙ on the SOHO plane of the sky during a complete solar rotation (from 1997 June 1 to 1997 June 28). The 28-days data sequence allows us to construct coronal off-limb Carrington maps of the resulting <span class="hlt">velocities</span> at different heliocentric distances to investigate the space and time evolution of the outflowing solar plasma. In addition, we performed a parameter space exploration in order to study the dependence of the derived outflow <span class="hlt">velocities</span> on the physical quantities characterising the Lyα emitting process in the corona. Our results are important in anticipation of the future science with the Metis instrument, selected to be part of the Solar Orbiter scientific payload. It was conceived to carry out near-sun coronagraphy, performing for the first time simultaneous imaging in polarised visible-light and ultraviolet H I Lyα line, so providing an unprecedented view of the solar wind acceleration region in the inner corona. The movie (see Sect. 4.2) is available at https://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMNG23A1376B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMNG23A1376B"><span>Possible high sonic <span class="hlt">velocity</span> due to the inclusion of gas bubbles in water</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Banno, T.; Mikada, H.; Goto, T.; Takekawa, J.</p> <p>2010-12-01</p> <p> equation when the liquid is supposed to be incompressible. We got the boundary integral equation from the Laplace equation and solved the boundary integral equation by the BEM. Then, we got the <span class="hlt">gradient</span> of the <span class="hlt">velocity</span> potential from the BEM. We used this <span class="hlt">gradient</span> to get time derivative of the <span class="hlt">velocity</span> potential from the Bernouii’s equation. And we used the second order Adams-Bashforth method to execute time integration of the <span class="hlt">velocity</span> potential. We conducted this scheme iteratively to calculate a bubble oscillation. At each time step, we input a pressure change as a sinusoidal wave. As a result, we <span class="hlt">observed</span> a bubble oscillation following the pressure frequency. We also evaluated the resonance frequency of a bubble by changing the pressure frequency. It showed a good agreement with the analytical solution described above. Our future work is to extend the calculation into plural bubbles condition. We expect that interaction between bubbles becomes strong and resonance frequency of bubbles becomes small when distance between bubbles becomes small.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29698471','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29698471"><span>Dose <span class="hlt">gradient</span> curve: A new tool for evaluating dose <span class="hlt">gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sung, KiHoon; Choi, Young Eun</p> <p>2018-01-01</p> <p>Stereotactic radiotherapy, which delivers an ablative high radiation dose to a target volume for maximum local tumor control, requires a rapid dose fall-off outside the target volume to prevent extensive damage to nearby normal tissue. Currently, there is no tool to comprehensively evaluate the dose <span class="hlt">gradient</span> near the target volume. We propose the dose <span class="hlt">gradient</span> curve (DGC) as a new tool to evaluate the quality of a treatment plan with respect to the dose fall-off characteristics. The average distance between two isodose surfaces was represented by the dose <span class="hlt">gradient</span> index (DGI) estimated by a simple equation using the volume and surface area of isodose levels. The surface area was calculated by mesh generation and surface triangulation. The DGC was defined as a plot of the DGI of each dose interval as a function of the dose. Two types of DGCs, differential and cumulative, were generated. The performance of the DGC was evaluated using stereotactic radiosurgery plans for virtual targets. Over the range of dose distributions, the dose <span class="hlt">gradient</span> of each dose interval was well-characterized by the DGC in an easily understandable graph format. Significant changes in the DGC were <span class="hlt">observed</span> reflecting the differences in planning situations and various prescription doses. The DGC is a rational method for visualizing the dose <span class="hlt">gradient</span> as the average distance between two isodose surfaces; the shorter the distance, the steeper the dose <span class="hlt">gradient</span>. By combining the DGC with the dose-volume histogram (DVH) in a single plot, the DGC can be utilized to evaluate not only the dose <span class="hlt">gradient</span> but also the target coverage in routine clinical practice.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5919624','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5919624"><span>Dose <span class="hlt">gradient</span> curve: A new tool for evaluating dose <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Choi, Young Eun</p> <p>2018-01-01</p> <p>Purpose Stereotactic radiotherapy, which delivers an ablative high radiation dose to a target volume for maximum local tumor control, requires a rapid dose fall-off outside the target volume to prevent extensive damage to nearby normal tissue. Currently, there is no tool to comprehensively evaluate the dose <span class="hlt">gradient</span> near the target volume. We propose the dose <span class="hlt">gradient</span> curve (DGC) as a new tool to evaluate the quality of a treatment plan with respect to the dose fall-off characteristics. Methods The average distance between two isodose surfaces was represented by the dose <span class="hlt">gradient</span> index (DGI) estimated by a simple equation using the volume and surface area of isodose levels. The surface area was calculated by mesh generation and surface triangulation. The DGC was defined as a plot of the DGI of each dose interval as a function of the dose. Two types of DGCs, differential and cumulative, were generated. The performance of the DGC was evaluated using stereotactic radiosurgery plans for virtual targets. Results Over the range of dose distributions, the dose <span class="hlt">gradient</span> of each dose interval was well-characterized by the DGC in an easily understandable graph format. Significant changes in the DGC were <span class="hlt">observed</span> reflecting the differences in planning situations and various prescription doses. Conclusions The DGC is a rational method for visualizing the dose <span class="hlt">gradient</span> as the average distance between two isodose surfaces; the shorter the distance, the steeper the dose <span class="hlt">gradient</span>. By combining the DGC with the dose-volume histogram (DVH) in a single plot, the DGC can be utilized to evaluate not only the dose <span class="hlt">gradient</span> but also the target coverage in routine clinical practice. PMID:29698471</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840031379&hterms=geophysique&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dgeophysique','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840031379&hterms=geophysique&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dgeophysique"><span>Dynamics of a surge <span class="hlt">observed</span> in the C IV and H alpha lines</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schmieder, B.; Mein, P.; Vial, J.-C.; Tandberg-Hanssen, E.</p> <p>1983-01-01</p> <p>Time sequences of a surge have been obtained in Active Region 2701 during a coordinated SMY program, on October 2nd, 1980, while the MSDP spectrograph operated in H-alpha at the Meudon Solar Tower and the UVSP spectrometer on SMM <span class="hlt">observed</span> in the 1548 A C IV resonance line. The cold (H-alpha) and hot (C IV) material follow the same channel, and the event lasts about 10 min in both lines. A good correlation is found between H-alpha and C IV <span class="hlt">velocities</span>; radial <span class="hlt">velocities</span> along the surge are in the range 40-60 km/s in both cases. The <span class="hlt">observations</span> are consistent with the hypothesis that a pressure <span class="hlt">gradient</span> drives the surge. The H-alpha data seem to indicate the presence of a shock wave in the chromosphere, while the C IV quantities (<span class="hlt">velocities</span>, accelerations) vary on a very short time scale. Their maxima occur at some locations which could be interpreted as 'pinched' zones.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/46250','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/46250"><span>Photographic guidance for selecting flow resistance coefficients in high-<span class="hlt">gradient</span> channels</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Steven E. Yochum; Francesco Comiti; Ellen Wohl; Gabrielle C. L. David; Luca Mao</p> <p>2014-01-01</p> <p>Photographic guidance is presented to assist with the estimation of Manning’s n and Darcy-Weisbach f in high-<span class="hlt">gradient</span> plane-bed, step-pool, and cascade channels. Reaches both with and without instream wood are included. These coefficients are necessary for the estimation of reachaverage <span class="hlt">velocity</span>, energy loss, and...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JGR...10611691M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JGR...10611691M"><span>A laboratory study of mean flow generation in rotating fluids by Reynolds stress <span class="hlt">gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGuinness, D. S.; Boyer, D. L.; Fernando, H. J. S.</p> <p>2001-06-01</p> <p>Laboratory experiments were conducted that demonstrate that a mean azimuthal flow can be produced by introducing Reynolds stress <span class="hlt">gradients</span> to a rotating fluid with zero initial mean flow. This mechanism may play a role in the generation of mean currents in coastal regions. The experiments entail the establishment of turbulence in a thin annular-shaped region centered within a cylindrical test cell through the use of a vertically oscillating grid. This region rests in a horizontal plane perpendicular to the vertical axis of the tank, and the entire system is placed on a turntable to simulate background rotation. Flow visualization techniques are used to depict qualitative features of the resulting flow field. Measurements of the mean and turbulent <span class="hlt">velocity</span> fields are performed using a two-component laser-Doppler velocimeter. The results show how rectified currents (mean flows) can be generated via Reynolds stress <span class="hlt">gradients</span> induced by periodic forcing of the grid. In the absence of background rotation, rectified flow is <span class="hlt">observed</span> in the radial and vertical directions only. The presence of background rotation tends to organize these motions in that the flow tends to move parallel to the turbulent source, i.e., in the azimuthal direction, with the source (strong turbulence) located to the right, facing downstream. The influence of rotation on the Reynolds stresses and their <span class="hlt">gradients</span> as well as on the ensuing mean flow is evaluated, and the <span class="hlt">observations</span> are examined by considering individual contributions of the terms in the Reynolds-averaged momentum equations.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17485673','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17485673"><span>Indoor seismology by probing the Earth's interior by using sound <span class="hlt">velocity</span> measurements at high pressures and temperatures.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Baosheng; Liebermann, Robert C</p> <p>2007-05-29</p> <p>The adiabatic bulk (K(S)) and shear (G) moduli of mantle materials at high pressure and temperature can be obtained directly by measuring compressional and shear wave <span class="hlt">velocities</span> in the laboratory with experimental techniques based on physical acoustics. We present the application of the current state-of-the-art experimental techniques by using ultrasonic interferometry in conjunction with synchrotron x radiation to study the elasticity of olivine and pyroxenes and their high-pressure phases. By using these updated thermoelasticity data for these phases, <span class="hlt">velocity</span> and density profiles for a pyrolite model are constructed and compared with radial seismic models. We conclude that pyrolite provides an adequate explanation of the major seismic discontinuities at 410- and 660-km depths, the <span class="hlt">gradient</span> in the transition zone, as well as the <span class="hlt">velocities</span> in the lower mantle, if the uncertainties in the modeling and the variations in different seismic models are considered. The characteristics of the seismic scaling factors in response to thermal anomalies suggest that anticorrelations between bulk sound and shear wave <span class="hlt">velocities</span>, as well as the large positive density anomalies <span class="hlt">observed</span> in the lower mantle, cannot be explained fully without invoking chemical variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36...91B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36...91B"><span>Characteristics of equatorial plasma bubbles <span class="hlt">observed</span> by TEC map based on ground-based GNSS receivers over South America</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barros, Diego; Takahashi, Hisao; Wrasse, Cristiano M.; Figueiredo, Cosme Alexandre O. B.</p> <p>2018-01-01</p> <p>A ground-based network of GNSS receivers has been used to monitor equatorial plasma bubbles (EPBs) by mapping the total electron content (TEC map). The large coverage of the TEC map allowed us to monitor several EPBs simultaneously and get characteristics of the dynamics, extension and longitudinal distributions of the EPBs from the onset time until their disappearance. These characteristics were obtained by using TEC map analysis and the keogram technique. TEC map databases analyzed were for the period between November 2012 and January 2016. The zonal drift <span class="hlt">velocities</span> of the EPBs showed a clear latitudinal <span class="hlt">gradient</span> varying from 123 m s-1 at the Equator to 65 m s-1 for 35° S latitude. Consequently, <span class="hlt">observed</span> EPBs are inclined against the geomagnetic field lines. Both zonal drift <span class="hlt">velocity</span> and the inclination of the EPBs were compared to the thermospheric neutral wind, which showed good agreement. Moreover, the large two-dimensional coverage of TEC maps allowed us to study periodic EPBs with a wide longitudinal distance. The averaged values <span class="hlt">observed</span> for the inter-bubble distances also presented a clear latitudinal <span class="hlt">gradient</span> varying from 920 km at the Equator to 640 km at 30° S. The latitudinal <span class="hlt">gradient</span> in the inter-bubble distances seems to be related to the difference in the zonal drift <span class="hlt">velocity</span> of the EPB from the Equator to middle latitudes and to the difference in the westward movement of the terminator. On several occasions, the distances reached more than 2000 km. Inter-bubble distances greater than 1000 km have not been reported in the literature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1355031-sound-velocity-density-magnesiowustites-implications-ultralow-velocity-zone-topography-sound-velocities-iron-rich-oxides','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1355031-sound-velocity-density-magnesiowustites-implications-ultralow-velocity-zone-topography-sound-velocities-iron-rich-oxides"><span>Sound <span class="hlt">velocity</span> and density of magnesiowüstites: Implications for ultralow-<span class="hlt">velocity</span> zone topography: Sound <span class="hlt">velocities</span> of Iron-rich Oxides</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wicks, June; Jackson, Jennifer M.; Sturhahn, Wolfgang</p> <p></p> <p>We explore the effect of Mg/Fe substitution on the sound <span class="hlt">velocities</span> of iron-rich (Mg 1 - xFe x)O, where x = 0.84, 0.94, and 1.0. Sound <span class="hlt">velocities</span> were determined using nuclear resonance inelastic X-ray scattering as a function of pressure, approaching those of the lowermost mantle. The systematics of cation substitution in the Fe-rich limit has the potential to play an important role in the interpretation of seismic <span class="hlt">observations</span> of the core-mantle boundary. By determining a relationship between sound <span class="hlt">velocity</span>, density, and composition of (Mg,Fe)O, this study explores the potential constraints on ultralow-<span class="hlt">velocity</span> zones at the core-mantle boundary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.477.2560B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.477.2560B"><span>M*/L <span class="hlt">gradients</span> driven by IMF variation: large impact on dynamical stellar mass estimates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bernardi, M.; Sheth, R. K.; Dominguez-Sanchez, H.; Fischer, J.-L.; Chae, K.-H.; Huertas-Company, M.; Shankar, F.</p> <p>2018-06-01</p> <p>Within a galaxy the stellar mass-to-light ratio ϒ* is not constant. Recent studies of spatially resolved kinematics of nearby early-type galaxies suggest that allowing for a variable initial mass function (IMF) returns significantly larger ϒ* <span class="hlt">gradients</span> than if the IMF is held fixed. We show that ignoring such IMF-driven ϒ* <span class="hlt">gradients</span> can have dramatic effect on dynamical (M_*^dyn), though stellar population (M_*^SP) based estimates of early-type galaxy stellar masses are also affected. This is because M_*^dyn is usually calibrated using the <span class="hlt">velocity</span> dispersion measured in the central regions (e.g. Re/8) where stars are expected to dominate the mass (i.e. the dark matter fraction is small). On the other hand, M_*^SP is often computed from larger apertures (e.g. using a mean ϒ* estimated from colours). If ϒ* is greater in the central regions, then ignoring the <span class="hlt">gradient</span> can overestimate M_*^dyn by as much as a factor of two for the most massive galaxies. Large ϒ*-<span class="hlt">gradients</span> have four main consequences: First, M_*^dyn cannot be estimated independently of stellar population synthesis models. Secondly, if there is a lower limit to ϒ* and <span class="hlt">gradients</span> are unknown, then requiring M_*^dyn=M_*^SP constrains them. Thirdly, if <span class="hlt">gradients</span> are stronger in more massive galaxies, then accounting for this reduces the slope of the correlation between M_*^dyn/M_*^SP of a galaxy with its <span class="hlt">velocity</span> dispersion. In particular, IMF-driven <span class="hlt">gradients</span> bring M_*^dyn and M_*^SP into agreement, not by shifting M_*^SP upwards by invoking constant bottom-heavy IMFs, as advocated by a number of recent studies, but by revising M_*^dyn estimates in the literature downwards. Fourthly, accounting for ϒ* <span class="hlt">gradients</span> changes the high-mass slope of the stellar mass function φ (M_*^dyn), and reduces the associated stellar mass density. These conclusions potentially impact estimates of the need for feedback and adiabatic contraction, so our results highlight the importance of measuring ϒ* <span class="hlt">gradients</span> in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G43B0934O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G43B0934O"><span>Dense <span class="hlt">Velocity</span> Field of Turkey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ozener, H.; Aktug, B.; Dogru, A.; Tasci, L.</p> <p>2017-12-01</p> <p>While the GNSS-based crustal deformation studies in Turkey date back to early 1990s, a homogenous <span class="hlt">velocity</span> field utilizing all the available data is still missing. Regional studies employing different site distributions, <span class="hlt">observation</span> plans, processing software and methodology not only create reference frame variations but also heterogeneous stochastic models. While the reference frame effect between different <span class="hlt">velocity</span> fields could easily be removed by estimating a set of rotations, the homogenization of the stochastic models of the individual <span class="hlt">velocity</span> fields requires a more detailed analysis. Using a rigorous Variance Component Estimation (VCE) methodology, we estimated the variance factors for each of the contributing <span class="hlt">velocity</span> fields and combined them into a single homogenous <span class="hlt">velocity</span> field covering whole Turkey. Results show that variance factors between <span class="hlt">velocity</span> fields including the survey mode and continuous <span class="hlt">observations</span> can vary a few orders of magnitude. In this study, we present the most complete <span class="hlt">velocity</span> field in Turkey rigorously combined from 20 individual <span class="hlt">velocity</span> fields including the 146 station CORS network and totally 1072 stations. In addition, three GPS campaigns were performed along the North Anatolian Fault and Aegean Region to fill the gap between existing <span class="hlt">velocity</span> fields. The homogenously combined new <span class="hlt">velocity</span> field is nearly complete in terms of geographic coverage, and will serve as the basis for further analyses such as the estimation of the deformation rates and the determination of the slip rates across main fault zones.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19810003848','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810003848"><span>Experimental analysis of the boundary layer transition with zero and positive pressure <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Arnal, D.; Jullen, J. C.; Michel, R.</p> <p>1980-01-01</p> <p>The influence of a positive pressure <span class="hlt">gradient</span> on the boundary layer transition is studied. The mean <span class="hlt">velocity</span> and turbulence profiles of four cases are examined. As the intensity of the pressure <span class="hlt">gradient</span> is increased, the Reynolds number of the transition onset and the length of the transition region are reduced. The Tollmein-Schlichting waves disturb the laminar regime; the amplification of these waves is in good agreement with the stability theory. The three dimensional deformation of the waves leads finally to the appearance of turbulence. In the case of zero pressure <span class="hlt">gradient</span>, the properties of the turbulent spots are studied by conditional sampling of the hot-wire signal; in the case of positive pressure <span class="hlt">gradient</span>, the turbulence appears in a progressive manner and the turbulent spots are much more difficult to characterize.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29532907','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29532907"><span>The Effect of Hydraulic <span class="hlt">Gradient</span> and Pattern of Conduit Systems on Tracing Tests: Bench-Scale Modeling.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mohammadi, Zargham; Gharaat, Mohammad Javad; Field, Malcolm</p> <p>2018-03-13</p> <p>Tracer breakthrough curves provide valuable information about the traced media, especially in inherently heterogeneous karst aquifers. In order to study the effect of variations in hydraulic <span class="hlt">gradient</span> and conduit systems on breakthrough curves, a bench scale karst model was constructed. The bench scale karst model contains both matrix and a conduit. Eight tracing tests were conducted under a wide range of hydraulic <span class="hlt">gradients</span> from 1 to greater than 5 for branchwork and network-conduit systems. Sampling points at varying distances from the injection point were utilized. Results demonstrate that mean tracer <span class="hlt">velocities</span>, tracer mass recovery and linear rising slope of the breakthrough curves were directly controlled by hydraulic <span class="hlt">gradient</span>. As hydraulic <span class="hlt">gradient</span> increased, both one half the time for peak concentration and one fifth the time for peak concentration decreased. The results demonstrate the variations in one half the time for peak concentration and one fifth the time for peak concentration of the descending limb for different sampling points under differing hydraulic <span class="hlt">gradients</span> are mainly controlled by the interactions of advection with dispersion. The results are discussed from three perspectives: different conduit systems, different hydraulic-<span class="hlt">gradient</span> conditions, and different sampling points. The research confirmed the undeniable role of hydrogeological setting (i.e., hydraulic <span class="hlt">gradient</span> and conduit system) on the shape of the breakthrough curve. The extracted parameters (mobile-fluid <span class="hlt">velocity</span>, tracer-mass recovery, linear rising limb, one half the time for peak concentration, and one fifth the time for peak concentration) allow for differentiating hydrogeological settings and enhance interpretations the tracing tests in karst aquifers. © 2018, National Ground Water Association.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120009904','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120009904"><span>Sodium <span class="hlt">Velocity</span> Maps on Mercury</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Potter, A. E.; Killen, R. M.</p> <p>2011-01-01</p> <p>The objective of the current work was to measure two-dimensional maps of sodium <span class="hlt">velocities</span> on the Mercury surface and examine the maps for evidence of sources or sinks of sodium on the surface. The McMath-Pierce Solar Telescope and the Stellar Spectrograph were used to measure Mercury spectra that were sampled at 7 milliAngstrom intervals. <span class="hlt">Observations</span> were made each day during the period October 5-9, 2010. The dawn terminator was in view during that time. The <span class="hlt">velocity</span> shift of the centroid of the Mercury emission line was measured relative to the solar sodium Fraunhofer line corrected for radial <span class="hlt">velocity</span> of the Earth. The difference between the <span class="hlt">observed</span> and calculated <span class="hlt">velocity</span> shift was taken to be the <span class="hlt">velocity</span> vector of the sodium relative to Earth. For each position of the spectrograph slit, a line of <span class="hlt">velocities</span> across the planet was measured. Then, the spectrograph slit was stepped over the surface of Mercury at 1 arc second intervals. The position of Mercury was stabilized by an adaptive optics system. The collection of lines were assembled into an images of surface reflection, sodium emission intensities, and Earthward <span class="hlt">velocities</span> over the surface of Mercury. The <span class="hlt">velocity</span> map shows patches of higher <span class="hlt">velocity</span> in the southern hemisphere, suggesting the existence of sodium sources there. The peak earthward <span class="hlt">velocity</span> occurs in the equatorial region, and extends to the terminator. Since this was a dawn terminator, this might be an indication of dawn evaporation of sodium. Leblanc et al. (2008) have published a <span class="hlt">velocity</span> map that is similar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMDI51A2659A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMDI51A2659A"><span>Comparison of Oceanic and Continental Lithosphere, Asthenosphere, and the LAB Through Shear <span class="hlt">Velocity</span> Inversion of Rayleigh Wave Data from the ALBACORE Amphibious Array in Southern California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Amodeo, K.; Rathnayaka, S.; Weeraratne, D. S.; Kohler, M. D.</p> <p>2016-12-01</p> <p>Continental and oceanic lithosphere, which form in different tectonic environments, are studied in a single amphibious seismic array across the Southern California continental margin. This provides a unique opportunity to directly compare oceanic and continental lithosphere, asthenosphere, and the LAB (Lithosphere-Asthenosphere Boundary) in a single data set. The complex history of the region, including spreading center subduction, block rotation, and Borderland extension, allows us to study limits in the rigidity and strength of the lithosphere. We study Rayleigh wave phase <span class="hlt">velocities</span> obtained from the ALBACORE (Asthenospheric and Lithospheric Broadband Architecture from the California Offshore Region Experiment) offshore seismic array project and invert for shear wave <span class="hlt">velocity</span> structure as a function of depth. We divide the study area into several regions: continent, inner Borderland, outer Borderland, and oceanic seafloor categorized by age. A unique starting Vs model is used for each case including layer thicknesses, densities, and P and S <span class="hlt">velocities</span> which predicts Rayleigh phase <span class="hlt">velocities</span> and are compared to <span class="hlt">observed</span> phase <span class="hlt">velocities</span> in each region. We solve for shear wave <span class="hlt">velocities</span> with the best fit between <span class="hlt">observed</span> and predicted phase <span class="hlt">velocity</span> data in a least square sense. Preliminary results indicate that lithospheric <span class="hlt">velocities</span> in the oceanic mantle are higher than the continental region by at least 2%. The LAB is <span class="hlt">observed</span> at 50 ± 20 km beneath 15-35 Ma oceanic seafloor. Asthenospheric low <span class="hlt">velocities</span> reach a minimum of 4.2 km/s in all regions, but have a steeper positive <span class="hlt">velocity</span> <span class="hlt">gradient</span> at the base of the oceanic asthenosphere compared to the continent. Seismic tomography images in two and three dimensions will be presented from each study region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSPO13C..05B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSPO13C..05B"><span>Retrieving Mesoscale Vertical <span class="hlt">Velocities</span> along the Antarctic Circumpolar Current from a Combination of Satellite and In Situ <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Buongiorno Nardelli, B.; Iudicone, D.; Cotroneo, Y.; Zambianchi, E.; Rio, M. H.</p> <p>2016-02-01</p> <p>In the framework of the Italian National Program on Antarctic Research (PNRA), an analysis of the mesoscale dynamics along the Antarctic Circumpolar Current has been carried out starting from a combination of satellite and in situ <span class="hlt">observations</span>. More specifically, state-of-the-art statistical techniques have been used to combine remotely-sensed sea surface temperature, salinity and absolute dynamical topography with in situ Argo data, providing mesoscale-resolving 3D tracers and geostrophic <span class="hlt">velocity</span> fields. The 3D reconstruction has been validated with independent data collected during PNRA surveys. These data are then used to diagnose the vertical exchanges in the Southern Ocean through a generalized version of the Omega equation. Intense vertical motion (O(100 m/day)) is found along the ACC, upstream/downstream of its meanders, and within mesoscale eddies, where multipolar vertical <span class="hlt">velocity</span> patterns are generally <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160010398','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160010398"><span>An X-Ray and Radio Study of the Varying Expansion <span class="hlt">Velocities</span> in Tycho's Supernova Remnant</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Williams, Brian J.; Chomiuk, Laura; Hewitt, John W.; Blondin, John M.; Borkowski, Kazimierz J.; Ghavamian, Parviz; Petre, Robert; Reynolds, Stephen P.</p> <p>2016-01-01</p> <p>We present newly obtained X-ray and radio <span class="hlt">observations</span> of Tycho's supernova remnant using Chandra and the Karl G. Jansky Very Large Array in 2015 and 2013/14, respectively. When combined with earlier epoch <span class="hlt">observations</span> by these instruments, we now have time baselines for expansion measurements of the remnant of 12-15 year in the X-rays and 30 year in the radio. The remnant's large angular size allows for proper motion measurements at many locations around the periphery of the blast wave. We find, consistent with earlier measurements, a clear <span class="hlt">gradient</span> in the expansion <span class="hlt">velocity</span> of the remnant, despite its round shape. The proper motions on the western and southwestern sides of the remnant are about a factor of two higher than those in the east and northeast. We showed in an earlier work that this is related to an offset of the explosion site from the geometric center of the remnant due to a density <span class="hlt">gradient</span> in the ISM, and using our refined measurements reported here, we find that this offset is approximately 23'' towards the northeast. An explosion center offset in such a circular remnant has implications for searches for progenitor companions in other remnants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFDD21003K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFDD21003K"><span>Application of the High <span class="hlt">Gradient</span> hydrodynamics code to simulations of a two-dimensional zero-pressure-<span class="hlt">gradient</span> turbulent boundary layer over a flat plate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kaiser, Bryan E.; Poroseva, Svetlana V.; Canfield, Jesse M.; Sauer, Jeremy A.; Linn, Rodman R.</p> <p>2013-11-01</p> <p>The High <span class="hlt">Gradient</span> hydrodynamics (HIGRAD) code is an atmospheric computational fluid dynamics code created by Los Alamos National Laboratory to accurately represent flows characterized by sharp <span class="hlt">gradients</span> in <span class="hlt">velocity</span>, concentration, and temperature. HIGRAD uses a fully compressible finite-volume formulation for explicit Large Eddy Simulation (LES) and features an advection scheme that is second-order accurate in time and space. In the current study, boundary conditions implemented in HIGRAD are varied to find those that better reproduce the reduced physics of a flat plate boundary layer to compare with complex physics of the atmospheric boundary layer. Numerical predictions are compared with available DNS, experimental, and LES data obtained by other researchers. High-order turbulence statistics are collected. The Reynolds number based on the free-stream <span class="hlt">velocity</span> and the momentum thickness is 120 at the inflow and the Mach number for the flow is 0.2. Results are compared at Reynolds numbers of 670 and 1410. A part of the material is based upon work supported by NASA under award NNX12AJ61A and by the Junior Faculty UNM-LANL Collaborative Research Grant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870033194&hterms=coulomb+law&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcoulomb%2Blaw','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870033194&hterms=coulomb+law&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcoulomb%2Blaw"><span>Beam-induced pressure <span class="hlt">gradients</span> in the early phase of proton-heated solar flares</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tamres, David H.; Canfield, Richard C.; Mcclymont, A. N.</p> <p>1986-01-01</p> <p>The pressure <span class="hlt">gradient</span> induced in a coronal loop by proton beam momentum deposition is calculated and compared with the thermal pressure <span class="hlt">gradient</span> arising from nonuniform deposition of beam energy; it is assumed that the transfer of momentum and energy from beam to target occurs via the Coulomb interaciton. Results are presented for both a low mean energy and a high mean energy proton beam injected at the loop apex and characterized by a power-law energy spectrum. The present treatment takes account of the breakdown of the cold target approximation for the low-energy proton beam in the corona, where the thermal speed of target electrons exceeds the beam speed. It is found that proton beam momentum deposition plays a potentially significant role in flare dynamics only in the low mean energy case and only in the corona, where it may dominate the acceleration of target material for as long as several tens of seconds. This conclusion suggest that the presence of low-energy nonthermal protons may be inferred from <span class="hlt">velocity</span>-sensitive coronal <span class="hlt">observations</span> in the early impulsive phase.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...837...88B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...837...88B"><span><span class="hlt">Velocity</span> Segregation and Systematic Biases In <span class="hlt">Velocity</span> Dispersion Estimates with the SPT-GMOS Spectroscopic Survey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bayliss, Matthew. B.; Zengo, Kyle; Ruel, Jonathan; Benson, Bradford A.; Bleem, Lindsey E.; Bocquet, Sebastian; Bulbul, Esra; Brodwin, Mark; Capasso, Raffaella; Chiu, I.-non; McDonald, Michael; Rapetti, David; Saro, Alex; Stalder, Brian; Stark, Antony A.; Strazzullo, Veronica; Stubbs, Christopher W.; Zenteno, Alfredo</p> <p>2017-03-01</p> <p>The <span class="hlt">velocity</span> distribution of galaxies in clusters is not universal; rather, galaxies are segregated according to their spectral type and relative luminosity. We examine the <span class="hlt">velocity</span> distributions of different populations of galaxies within 89 Sunyaev Zel’dovich (SZ) selected galaxy clusters spanning 0.28< z< 1.08. Our sample is primarily draw from the SPT-GMOS spectroscopic survey, supplemented by additional published spectroscopy, resulting in a final spectroscopic sample of 4148 galaxy spectra—2868 cluster members. The <span class="hlt">velocity</span> dispersion of star-forming cluster galaxies is 17 ± 4% greater than that of passive cluster galaxies, and the <span class="hlt">velocity</span> dispersion of bright (m< {m}* -0.5) cluster galaxies is 11 ± 4% lower than the <span class="hlt">velocity</span> dispersion of our total member population. We find good agreement with simulations regarding the shape of the relationship between the measured <span class="hlt">velocity</span> dispersion and the fraction of passive versus star-forming galaxies used to measure it, but we find a small offset between this relationship as measured in data and simulations, which suggests that our dispersions are systematically low by as much as 3% relative to simulations. We argue that this offset could be interpreted as a measurement of the effective <span class="hlt">velocity</span> bias that describes the ratio of our <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersions and the intrinsic <span class="hlt">velocity</span> dispersion of dark matter particles in a published simulation result. Measuring <span class="hlt">velocity</span> bias in this way suggests that large spectroscopic surveys can improve dispersion-based mass-<span class="hlt">observable</span> scaling relations for cosmology even in the face of <span class="hlt">velocity</span> biases, by quantifying and ultimately calibrating them out.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870029355&hterms=egg&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Degg','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870029355&hterms=egg&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Degg"><span>Aperture-synthesis <span class="hlt">observations</span> of carbon monoxide in the Egg Nebula</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heiligman, G. M.; Berge, G. L.; Claussen, M. J.; Leighton, R. B.; Lo, K. Y.; Masson, C. R.; Moffet, A. T.; Phillips, T. G.; Sargent, A. I.; Wannier, P. G.</p> <p>1986-01-01</p> <p><span class="hlt">Observations</span> of the 2.6-mm CO emission of the bipolar nebular CRL 2688, obtained with resolution 7 arcsec using the mm-wave interferometer at Owens Valley during December 1982-June 1983, are reported. The emission of a 10 x 15-arcsec core, centered on the optical reflection nebula and probably surrounded by a large cloud of cooler gas, is found to have a main-axis <span class="hlt">velocity</span> <span class="hlt">gradient</span> of 3 km/s arcsec and an excitation temperature of about 70 K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18185672','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18185672"><span><span class="hlt">Observation</span> of a single-beam <span class="hlt">gradient</span>-force optical trap for dielectric particles in air.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Omori, R; Kobayashi, T; Suzuki, A</p> <p>1997-06-01</p> <p>A single-beam <span class="hlt">gradient</span>-force optical trap for dielectric particles, which relies solely on the radiation pressure force of a TEM(00)-mode laser light, is demonstrated in air for what is believed to be the first time. It was <span class="hlt">observed</span> that micrometer-sized glass spheres with a refractive index of n=1.45 remained trapped in the focus region for more than 30 min, and we could transfer them three dimensionally by moving the beam focus and the microscope stage. A laser power of ~40 mW was sufficient to trap a 5- microm -diameter glass sphere. The present method has several distinct advantages over the conventional optical levitation method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1984/0733/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1984/0733/report.pdf"><span><span class="hlt">Velocity</span> profile, water-surface slope, and bed-material size for selected streams in Colorado</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Marchand, J.P.; Jarrett, R.D.; Jones, L.L.</p> <p>1984-01-01</p> <p>Existing methods for determining the mean <span class="hlt">velocity</span> in a vertical sampling section do not address the conditions present in high-<span class="hlt">gradient</span>, shallow-depth streams common to mountainous regions such as Colorado. The report presents <span class="hlt">velocity</span>-profile data that were collected for 11 streamflow-gaging stations in Colorado using both a standard Price type AA current meter and a prototype Price Model PAA current meter. Computational results are compiled that will enable mean <span class="hlt">velocities</span> calculated from measurements by the two current meters to be compared with each other and with existing methods for determining mean <span class="hlt">velocity</span>. Water-surface slope, bed-material size, and flow-characteristic data for the 11 sites studied also are presented. (USGS)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850026762','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850026762"><span>Solar activity beyond the disk and variations of the cosmic ray <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Belov, A. V.; Dorman, L. I.; Eroshenko, E. A.; Ishkov, V. N.; Oleneva, V. A.</p> <p>1985-01-01</p> <p>Part of galactic cosmic rays (CR) <span class="hlt">observed</span> near the Earth and on the Earth come from beyond-disk regions of circumsolar space. But CR of those energies which undergo substantial modulation cover too large a path across the lines of force of the interplanetary magnetic field (IMF) in order that they could provide an effective transfer of information about beyond-disk solar activity. And if it is still possible, the most probable channel for transferring such information must be a neutral layer of heliomagnetosphere in which the transverse CR transport is facilitated by their drift in an inhomogeneous magnetic field. A simple diffusion model for an expected CR variation in a neutral layer near the Earth is discussed. It is of importance that variations of the CR <span class="hlt">gradient</span> are not at all always accompanied by considerable variations of IMF and solar wind <span class="hlt">velocity</span> at the point of <span class="hlt">observation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ChJME..25..715A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ChJME..25..715A"><span>Calculation and analysis of <span class="hlt">velocity</span> and viscous drag in an artery with a periodic pressure <span class="hlt">gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alizadeh, M.; Seyedpour, S. M.; Mozafari, V.; Babazadeh, Shayan S.</p> <p>2012-07-01</p> <p>Blood as a fluid that human and other living creatures are dependent on has been always considered by scientists and researchers. Any changes in blood pressure and its normal <span class="hlt">velocity</span> can be a sign of a disease. Whatever significant in blood fluid's mechanics is Constitutive equations and finding some relations for analysis and description of drag, <span class="hlt">velocity</span> and periodic blood pressure in vessels. In this paper, by considering available experimental quantities, for blood pressure and <span class="hlt">velocity</span> in periodic time of a thigh artery of a living dog, at first it is written into Fourier series, then by solving Navier-Stokes equations, a relation for curve drawing of vessel blood pressure with rigid wall is obtained. Likewise in another part of this paper, vessel wall is taken in to consideration that vessel wall is elastic and its pressure and <span class="hlt">velocity</span> are written into complex Fourier series. In this case, by solving Navier-Stokes equations, some relations for blood <span class="hlt">velocity</span>, viscous drag on vessel wall and blood pressure are obtained. In this study by noting that vessel diameter is almost is large (3.7 mm), and blood is considered as a Newtonian fluid. Finally, available experimental quantities of pressure with obtained curve of solving Navier-Stokes equations are compared. In blood analysis in rigid vessel, existence of 48% variance in pressure curve systole peak caused vessel blood flow analysis with elastic wall, results in new relations for blood flow description. The Resultant curve is obtained from new relations holding 10% variance in systole peak.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JPhCS.362a2009D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JPhCS.362a2009D"><span><span class="hlt">Velocity</span> Inversion In Cylindrical Couette Gas Flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dongari, Nishanth; Barber, Robert W.; Emerson, David R.; Zhang, Yonghao; Reese, Jason M.</p> <p>2012-05-01</p> <p>We investigate a power-law probability distribution function to describe the mean free path of rarefied gas molecules in non-planar geometries. A new curvature-dependent model is derived by taking into account the boundary-limiting effects on the molecular mean free path for surfaces with both convex and concave curvatures. In comparison to a planar wall, we find that the mean free path for a convex surface is higher at the wall and exhibits a sharper <span class="hlt">gradient</span> within the Knudsen layer. In contrast, a concave wall exhibits a lower mean free path near the surface and the <span class="hlt">gradients</span> in the Knudsen layer are shallower. The Navier-Stokes constitutive relations and <span class="hlt">velocity</span>-slip boundary conditions are modified based on a power-law scaling to describe the mean free path, in accordance with the kinetic theory of gases, i.e. transport properties can be described in terms of the mean free path. <span class="hlt">Velocity</span> profiles for isothermal cylindrical Couette flow are obtained using the power-law model. We demonstrate that our model is more accurate than the classical slip solution, especially in the transition regime, and we are able to capture important non-linear trends associated with the non-equilibrium physics of the Knudsen layer. In addition, we establish a new criterion for the critical accommodation coefficient that leads to the non-intuitive phenomena of <span class="hlt">velocity</span>-inversion. Our results are compared with conventional hydrodynamic models and direct simulation Monte Carlo data. The power-law model predicts that the critical accommodation coefficient is significantly lower than that calculated using the classical slip solution and is in good agreement with available DSMC data. Our proposed constitutive scaling for non-planar surfaces is based on simple physical arguments and can be readily implemented in conventional fluid dynamics codes for arbitrary geometric configurations.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C43B0805B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C43B0805B"><span>High-resolution, terrestrial radar <span class="hlt">velocity</span> <span class="hlt">observations</span> and model results reveal a strong bed at stable, tidewater Rink Isbræ, West Greenland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bartholomaus, T. C.; Walker, R. T.; Stearns, L. A.; Fahnestock, M. A.; Cassotto, R.; Catania, G. A.; Felikson, D.; Fried, M.; Sutherland, D.; Nash, J. D.; Shroyer, E.</p> <p>2015-12-01</p> <p>At tidewater Rink Isbræ, on the central west coast of Greenland, satellite <span class="hlt">observations</span> reveal that glacier <span class="hlt">velocities</span> and terminus positions have remained stable, while the lowest 25 km have thinned 30 m since 1985. Over this same time period, other tidewater glaciers in central west Greenland have retreated, thinned and accelerated. Here we present field <span class="hlt">observations</span> and model results to show that the flow of Rink Isbræ is resisted by unusually high basal shear stresses. Terrestrial radar interferometry (TRI) <span class="hlt">observations</span> over 9 days in summer 2014 demonstrate weak <span class="hlt">velocity</span> response to 4 km wide, full thickness calving events. <span class="hlt">Velocities</span> at the terminus change by +/- 10% in response to rising and falling tides within a partial-width, 2.5-km-long floating ice tongue; however these tidal perturbations damp out within 2 km of the grounding line. Inversions for basal shear stress and force balance analyses together show that basal shear stresses in excess of 300 kPa support the majority of the driving stress at thick, steep Rink Isbræ. These <span class="hlt">observational</span> and modeling results tell a consistent story in which a strong bed may limit the unstable tidewater glacier retreats <span class="hlt">observed</span> elsewhere. Rink Isbræ has an erosion resistant quartzite bed with low fracture density. We hypothesize that this geology may play a major role in the bed strength.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGeod.tmp...24L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGeod.tmp...24L"><span>Refined discrete and empirical horizontal <span class="hlt">gradients</span> in VLBI analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Landskron, Daniel; Böhm, Johannes</p> <p>2018-02-01</p> <p>Missing or incorrect consideration of azimuthal asymmetry of troposphere delays is a considerable error source in space geodetic techniques such as Global Navigation Satellite Systems (GNSS) or Very Long Baseline Interferometry (VLBI). So-called horizontal troposphere <span class="hlt">gradients</span> are generally utilized for modeling such azimuthal variations and are particularly required for <span class="hlt">observations</span> at low elevation angles. Apart from estimating the <span class="hlt">gradients</span> within the data analysis, which has become common practice in space geodetic techniques, there is also the possibility to determine the <span class="hlt">gradients</span> beforehand from different data sources than the actual <span class="hlt">observations</span>. Using ray-tracing through Numerical Weather Models (NWMs), we determined discrete <span class="hlt">gradient</span> values referred to as GRAD for VLBI <span class="hlt">observations</span>, based on the standard <span class="hlt">gradient</span> model by Chen and Herring (J Geophys Res 102(B9):20489-20502, 1997. https://doi.org/10.1029/97JB01739) and also for new, higher-order <span class="hlt">gradient</span> models. These <span class="hlt">gradients</span> are produced on the same data basis as the Vienna Mapping Functions 3 (VMF3) (Landskron and Böhm in J Geod, 2017.https://doi.org/10.1007/s00190-017-1066-2), so they can also be regarded as the VMF3 <span class="hlt">gradients</span> as they are fully consistent with each other. From VLBI analyses of the Vienna VLBI and Satellite Software (VieVS), it becomes evident that baseline length repeatabilities (BLRs) are improved on average by 5% when using a priori <span class="hlt">gradients</span> GRAD instead of estimating the <span class="hlt">gradients</span>. The reason for this improvement is that the <span class="hlt">gradient</span> estimation yields poor results for VLBI sessions with a small number of <span class="hlt">observations</span>, while the GRAD a priori <span class="hlt">gradients</span> are unaffected from this. We also developed a new empirical <span class="hlt">gradient</span> model applicable for any time and location on Earth, which is included in the Global Pressure and Temperature 3 (GPT3) model. Although being able to describe only the systematic component of azimuthal asymmetry and no short-term variations at all, even these</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1345625-velocity-segregation-systematic-biases-velocity-dispersion-estimates-spt-gmos-spectroscopic-survey','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1345625-velocity-segregation-systematic-biases-velocity-dispersion-estimates-spt-gmos-spectroscopic-survey"><span><span class="hlt">Velocity</span> segregation and systematic biases in <span class="hlt">velocity</span> dispersion estimates with the SPT-GMOS spectroscopic survey</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Bayliss, Matthew. B.; Zengo, Kyle; Ruel, Jonathan; ...</p> <p>2017-03-07</p> <p>The <span class="hlt">velocity</span> distribution of galaxies in clusters is not universal; rather, galaxies are segregated according to their spectral type and relative luminosity. We examine the <span class="hlt">velocity</span> distributions of different populations of galaxies within 89 Sunyaev Zel'dovich (SZ) selected galaxy clusters spanningmore » $ 0.28 < z < 1.08$. Our sample is primarily draw from the SPT-GMOS spectroscopic survey, supplemented by additional published spectroscopy, resulting in a final spectroscopic sample of 4148 galaxy spectra---2868 cluster members. The <span class="hlt">velocity</span> dispersion of star-forming cluster galaxies is $$17\\pm4$$% greater than that of passive cluster galaxies, and the <span class="hlt">velocity</span> dispersion of bright ($$m < m^{*}-0.5$$) cluster galaxies is $$11\\pm4$$% lower than the <span class="hlt">velocity</span> dispersion of our total member population. We find good agreement with simulations regarding the shape of the relationship between the measured <span class="hlt">velocity</span> dispersion and the fraction of passive vs. star-forming galaxies used to measure it, but we find a small offset between this relationship as measured in data and simulations in which suggests that our dispersions are systematically low by as much as 3\\% relative to simulations. We argue that this offset could be interpreted as a measurement of the effective <span class="hlt">velocity</span> bias that describes the ratio of our <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersions and the intrinsic <span class="hlt">velocity</span> dispersion of dark matter particles in a published simulation result. Here, by measuring <span class="hlt">velocity</span> bias in this way suggests that large spectroscopic surveys can improve dispersion-based mass-<span class="hlt">observable</span> scaling relations for cosmology even in the face of <span class="hlt">velocity</span> biases, by quantifying and ultimately calibrating them out.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1345625','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1345625"><span><span class="hlt">Velocity</span> segregation and systematic biases in <span class="hlt">velocity</span> dispersion estimates with the SPT-GMOS spectroscopic survey</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bayliss, Matthew. B.; Zengo, Kyle; Ruel, Jonathan</p> <p></p> <p>The <span class="hlt">velocity</span> distribution of galaxies in clusters is not universal; rather, galaxies are segregated according to their spectral type and relative luminosity. We examine the <span class="hlt">velocity</span> distributions of different populations of galaxies within 89 Sunyaev Zel'dovich (SZ) selected galaxy clusters spanningmore » $ 0.28 < z < 1.08$. Our sample is primarily draw from the SPT-GMOS spectroscopic survey, supplemented by additional published spectroscopy, resulting in a final spectroscopic sample of 4148 galaxy spectra---2868 cluster members. The <span class="hlt">velocity</span> dispersion of star-forming cluster galaxies is $$17\\pm4$$% greater than that of passive cluster galaxies, and the <span class="hlt">velocity</span> dispersion of bright ($$m < m^{*}-0.5$$) cluster galaxies is $$11\\pm4$$% lower than the <span class="hlt">velocity</span> dispersion of our total member population. We find good agreement with simulations regarding the shape of the relationship between the measured <span class="hlt">velocity</span> dispersion and the fraction of passive vs. star-forming galaxies used to measure it, but we find a small offset between this relationship as measured in data and simulations in which suggests that our dispersions are systematically low by as much as 3\\% relative to simulations. We argue that this offset could be interpreted as a measurement of the effective <span class="hlt">velocity</span> bias that describes the ratio of our <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersions and the intrinsic <span class="hlt">velocity</span> dispersion of dark matter particles in a published simulation result. Here, by measuring <span class="hlt">velocity</span> bias in this way suggests that large spectroscopic surveys can improve dispersion-based mass-<span class="hlt">observable</span> scaling relations for cosmology even in the face of <span class="hlt">velocity</span> biases, by quantifying and ultimately calibrating them out.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900043855&hterms=Barry+Woods&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DBarry%2BWoods','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900043855&hterms=Barry+Woods&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DBarry%2BWoods"><span>Luminosity-<span class="hlt">velocity</span> diagrams for Virgo Cluster spirals. I - Inner rotation curves</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Woods, David; Fahlman, Gregory G.; Madore, Barry F.</p> <p>1990-01-01</p> <p>Optical rotation curves are presented for the innermost portions of nine spiral galaxies in the Virgo Cluster. The emission-line (H-alpha and forbidden N II) <span class="hlt">velocity</span> data are to be used in combination with new CCD photometry to construct luminosity-<span class="hlt">velocity</span> diagrams, in a continuing investigation of an apparent initial linear branch and its potential as a distance indicator. Compared to recent H I data, the present optical rotation curves generally show systematically steeper inner <span class="hlt">gradients</span>. This effect is ascribed to the poorer resolution of the H I data and/or to holes in the gas distribution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSAH12A..08T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSAH12A..08T"><span>Net community calcification and production rates from Palmyra Atoll using a boundary layer <span class="hlt">gradient</span> flux approach</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeshita, Y.; McGillis, W. R.; Martz, T. R.; Price, N.; Smith, J.; Donham, E. M.</p> <p>2016-02-01</p> <p>Coral reefs are a highly dynamic system, where large variability in environmental conditions (e.g. pH) occurs on timescales of minutes to hours. Yet, techniques that are capable of monitoring reef calcification rates without artificial confinement on the same frequency are scarce. Here, we present a 2 week time series of sub-hourly, in situ benthic net community production (Pnet) and net community calcification (Gnet) rates from a reef terrace at Palmyra Atoll using the Benthic Ecosystem and Acidification Monitoring System (BEAMS). The net metabolism rates reported here are measured under natural conditions, without any alterations to the environment (e.g. light, flow, pH). The BEAMS measures the chemical <span class="hlt">gradient</span> and the current <span class="hlt">velocity</span> profile in the benthic boundary layer using autonomous sensors to calculate the chemical flux from the benthos. The O2 and total alkalinity (TA) fluxes were used to calculate Pnet and Gnet, respectively; TA <span class="hlt">gradients</span> were calculated from pH and O2 measurements. Gnet can be constrained to better than 3 mmol CaCO3 m-2 hr-1 using this approach, based on three simultaneous BEAMS deployments. A clear diel cycle of Gnet was <span class="hlt">observed</span>, where the peak day time Gnet and average nighttime Gnet were 14 and 1 mmol CaCO3 m-2 hr-1, respectively. Integrated daily Gnet ranged from 76 to 219 mmol CaCO3 m-2 d-1, with an average of 107 ± 14 mmol CaCO3 m-2 d-1. Light had the strongest control over Gnet, with current <span class="hlt">velocity</span> having a smaller yet noticeable effect. During the deployment, pH varied by 0.16 (ranged between 7.92 and 8.08), and a significant positive relationship was <span class="hlt">observed</span> between pH and Gnet. However, pH was also positively correlated with current <span class="hlt">velocity</span> and Pnet, making it difficult to determine if natural variability in pH was significantly affecting Gnet on the timescale of days to weeks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140005825','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140005825"><span>Sudden Intensity Increases and Radial <span class="hlt">Gradient</span> Changes of Cosmic Ray Mev Electrons and Protons <span class="hlt">Observed</span> at Voyager 1 Beyond 111 AU in the Heliosheath</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Webber, W. R.; Mcdonald, F. B.; Cummings, A. C.; Stone, E. C.; Heikkila, B.; Lal, N.</p> <p>2012-01-01</p> <p>Voyager 1 has entered regions of different propagation conditions for energetic cosmic rays in the outer heliosheathat a distance of about 111 AU from the Sun. The low energy 614 MeV galactic electron intensity increased by 20over a time period 10 days and the electron radial intensity <span class="hlt">gradient</span> abruptly decreased from 19AU to 8AU at2009.7 at a radial distance of 111.2 AU. At about 2011.2 at a distance of 116.6 AU a second abrupt intensity increase of25 was <span class="hlt">observed</span> for electrons. After the second sudden electron increase the radial intensity <span class="hlt">gradient</span> increased to18AU. This large positive <span class="hlt">gradient</span> and the 13 day periodic variations of 200 MeV particles <span class="hlt">observed</span> near theend of 2011 indicate that V1 is still within the overall heliospheric modulating region. The implications of these resultsregarding the proximity of the heliopause are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PASJ...70....9H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PASJ...70....9H"><span>Atmospheric gas dynamics in the Perseus cluster <span class="hlt">observed</span> with Hitomi</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hitomi Collaboration; Aharonian, Felix; Akamatsu, Hiroki; Akimoto, Fumie; Allen, Steven W.; Angelini, Lorella; Audard, Marc; Awaki, Hisamitsu; Axelsson, Magnus; Bamba, Aya; Bautz, Marshall W.; Blandford, Roger; Brenneman, Laura W.; Brown, Gregory V.; Bulbul, Esra; Cackett, Edward M.; Canning, Rebecca E. A.; Chernyakova, Maria; Chiao, Meng P.; Coppi, Paolo S.; Costantini, Elisa; de Plaa, Jelle; de Vries, Cor P.; den Herder, Jan-Willem; Done, Chris; Dotani, Tadayasu; Ebisawa, Ken; Eckart, Megan E.; Enoto, Teruaki; Ezoe, Yuichiro; Fabian, Andrew C.; Ferrigno, Carlo; Foster, Adam R.; Fujimoto, Ryuichi; Fukazawa, Yasushi; Furuzawa, Akihiro; Galeazzi, Massimiliano; Gallo, Luigi C.; Gandhi, Poshak; Giustini, Margherita; Goldwurm, Andrea; Gu, Liyi; Guainazzi, Matteo; Haba, Yoshito; Hagino, Kouichi; Hamaguchi, Kenji; Harrus, Ilana M.; Hatsukade, Isamu; Hayashi, Katsuhiro; Hayashi, Takayuki; Hayashi, Tasuku; Hayashida, Kiyoshi; Hiraga, Junko S.; Hornschemeier, Ann; Hoshino, Akio; Hughes, John P.; Ichinohe, Yuto; Iizuka, Ryo; Inoue, Hajime; Inoue, Shota; Inoue, Yoshiyuki; Ishida, Manabu; Ishikawa, Kumi; Ishisaki, Yoshitaka; Iwai, Masachika; Kaastra, Jelle; Kallman, Tim; Kamae, Tsuneyoshi; Kataoka, Jun; Katsuda, Satoru; Kawai, Nobuyuki; Kelley, Richard L.; Kilbourne, Caroline A.; Kitaguchi, Takao; Kitamoto, Shunji; Kitayama, Tetsu; Kohmura, Takayoshi; Kokubun, Motohide; Koyama, Katsuji; Koyama, Shu; Kretschmar, Peter; Krimm, Hans A.; Kubota, Aya; Kunieda, Hideyo; Laurent, Philippe; Lee, Shiu-Hang; Leutenegger, Maurice A.; Limousin, Olivier; Loewenstein, Michael; Long, Knox S.; Lumb, David; Madejski, Greg; Maeda, Yoshitomo; Maier, Daniel; Makishima, Kazuo; Markevitch, Maxim; Matsumoto, Hironori; Matsushita, Kyoko; McCammon, Dan; McNamara, Brian R.; Mehdipour, Missagh; Miller, Eric D.; Miller, Jon M.; Mineshige, Shin; Mitsuda, Kazuhisa; Mitsuishi, Ikuyuki; Miyazawa, Takuya; Mizuno, Tsunefumi; Mori, Hideyuki; Mori, Koji; Mukai, Koji; Murakami, Hiroshi; Mushotzky, Richard F.; Nakagawa, Takao; Nakajima, Hiroshi; Nakamori, Takeshi; Nakashima, Shinya; Nakazawa, Kazuhiro; Nobukawa, Kumiko K.; Nobukawa, Masayoshi; Noda, Hirofumi; Odaka, Hirokazu; Ohashi, Takaya; Ohno, Masanori; Okajima, Takashi; Ota, Naomi; Ozaki, Masanobu; Paerels, Frits; Paltani, Stéphane; Petre, Robert; Pinto, Ciro; Porter, Frederick S.; Pottschmidt, Katja; Reynolds, Christopher S.; Safi-Harb, Samar; Saito, Shinya; Sakai, Kazuhiro; Sasaki, Toru; Sato, Goro; Sato, Kosuke; Sato, Rie; Sawada, Makoto; Schartel, Norbert; Serlemtsos, Peter J.; Seta, Hiromi; Shidatsu, Megumi; Simionescu, Aurora; Smith, Randall K.; Soong, Yang; Stawarz, Łukasz; Sugawara, Yasuharu; Sugita, Satoshi; Szymkowiak, Andrew; Tajima, Hiroyasu; Takahashi, Hiromitsu; Takahashi, Tadayuki; Takeda, Shin'ichiro; Takei, Yoh; Tamagawa, Toru; Tamura, Takayuki; Tanaka, Keigo; Tanaka, Takaaki; Tanaka, Yasuo; Tanaka, Yasuyuki T.; Tashiro, Makoto S.; Tawara, Yuzuru; Terada, Yukikatsu; Terashima, Yuichi; Tombesi, Francesco; Tomida, Hiroshi; Tsuboi, Yohko; Tsujimoto, Masahiro; Tsunemi, Hiroshi; Tsuru, Takeshi Go; Uchida, Hiroyuki; Uchiyama, Hideki; Uchiyama, Yasunobu; Ueda, Shutaro; Ueda, Yoshihiro; Uno, Shin'ichiro; Urry, C. Megan; Ursino, Eugenio; Wang, Qian H. S.; Watanabe, Shin; Werner, Norbert; Wilkins, Dan R.; Williams, Brian J.; Yamada, Shinya; Yamaguchi, Hiroya; Yamaoka, Kazutaka; Yamasaki, Noriko Y.; Yamauchi, Makoto; Yamauchi, Shigeo; Yaqoob, Tahir; Yatsu, Yoichi; Yonetoku, Daisuke; Zhuravleva, Irina; Zoghbi, Abderahmen</p> <p>2018-03-01</p> <p>Extending the earlier measurements reported in Hitomi collaboration (2016, Nature, 535, 117), we examine the atmospheric gas motions within the central 100 kpc of the Perseus cluster using <span class="hlt">observations</span> obtained with the Hitomi satellite. After correcting for the point spread function of the telescope and using optically thin emission lines, we find that the line-of-sight <span class="hlt">velocity</span> dispersion of the hot gas is remarkably low and mostly uniform. The <span class="hlt">velocity</span> dispersion reaches a maxima of approximately 200 km s-1 toward the central active galactic nucleus (AGN) and toward the AGN inflated northwestern "ghost" bubble. Elsewhere within the <span class="hlt">observed</span> region, the <span class="hlt">velocity</span> dispersion appears constant around 100 km s-1. We also detect a <span class="hlt">velocity</span> <span class="hlt">gradient</span> with a 100 km s-1 amplitude across the cluster core, consistent with large-scale sloshing of the core gas. If the <span class="hlt">observed</span> gas motions are isotropic, the kinetic pressure support is less than 10% of the thermal pressure support in the cluster core. The well-resolved, optically thin emission lines have Gaussian shapes, indicating that the turbulent driving scale is likely below 100 kpc, which is consistent with the size of the AGN jet inflated bubbles. We also report the first measurement of the ion temperature in the intracluster medium, which we find to be consistent with the electron temperature. In addition, we present a new measurement of the redshift of the brightest cluster galaxy NGC 1275.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17046778','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17046778"><span>Estimation of the zeta potential and the dielectric constant using <span class="hlt">velocity</span> measurements in the electroosmotic flows.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, H M; Hong, S M</p> <p>2006-12-15</p> <p>In this paper we develop a method for the determination of the zeta potential zeta and the dielectric constant epsilon by exploiting <span class="hlt">velocity</span> measurements of the electroosmotic flow in microchannels. The inverse problem is solved through the minimization of a performance function utilizing the conjugate <span class="hlt">gradient</span> method. The present method is found to estimate zeta and epsilon with reasonable accuracy even with noisy <span class="hlt">velocity</span> measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P53B2651P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P53B2651P"><span>Radial-<span class="hlt">Velocity</span> Signatures of Magnetic Features on the Sun <span class="hlt">Observed</span> as a Star</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Palumbo, M. L., III; Haywood, R. D.; Saar, S. H.; Dupree, A. K.; Milbourne, T. W.</p> <p>2017-12-01</p> <p>In recent years, the search for Earth-mass planets using radial-<span class="hlt">velocity</span> measurements has become increasingly limited by signals arising from stellar activity. Individual magnetic features induce localized changes in intensity and <span class="hlt">velocity</span>, which combine to change the apparent radial <span class="hlt">velocity</span> of the star. Therefore it is critical to identify an indicator of activity-driven radial-<span class="hlt">velocity</span> variations on the timescale of stellar rotation periods. We use 617.3 nm photospheric filtergrams, magnetograms, and dopplergrams from SDO/HMI and 170.0 nm chromospheric filtergrams from AIA to identify magnetically-driven solar features and reconstruct the integrated solar radial <span class="hlt">velocity</span> with six samples per day over the course of 2014. Breaking the solar image up into regions of umbrae, penumbrae, quiet Sun, network, and plages, we find a distinct variation in the center-to-limb intensity-weighted <span class="hlt">velocity</span> for each region. In agreement with past studies, we find that the suppression of convective blueshift is dominated by plages and network, rather than dark photospheric features. In the future, this work will be highly useful for identifying indicators which correlate with rotationally modulated radial-<span class="hlt">velocity</span> variations. This will allow us to break the activity barrier that currently precludes the precise characterization of exoplanet properties at the lowest masses. This work was supported by the NSF-REU solar physics program at SAO, grant number AGS-1560313. This work was performed in part under contract with the California Institute of Technology (Caltech)/Jet Propulsion Laboratory (JPL) funded by NASA through the Sagan Fellowship Program executed by the NASA Exoplanet Science Institute.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.466.4780M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.466.4780M"><span>Why do high-redshift galaxies show diverse gas-phase metallicity <span class="hlt">gradients</span>?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Xiangcheng; Hopkins, Philip F.; Feldmann, Robert; Torrey, Paul; Faucher-Giguère, Claude-André; Kereš, Dušan</p> <p>2017-04-01</p> <p>Recent spatially resolved <span class="hlt">observations</span> of galaxies at z ˜ 0.6-3 reveal that high-redshift galaxies show complex kinematics and a broad distribution of gas-phase metallicity <span class="hlt">gradients</span>. To understand these results, we use a suite of high-resolution cosmological zoom-in simulations from the Feedback in Realistic Environments project, which include physically motivated models of the multiphase interstellar medium, star formation and stellar feedback. Our simulations reproduce the <span class="hlt">observed</span> diversity of kinematic properties and metallicity <span class="hlt">gradients</span>, broadly consistent with <span class="hlt">observations</span> at z ˜ 0-3. Strong negative metallicity <span class="hlt">gradients</span> only appear in galaxies with a rotating disc, but not all rotationally supported galaxies have significant <span class="hlt">gradients</span>. Strongly perturbed galaxies with little rotation always have flat <span class="hlt">gradients</span>. The kinematic properties and metallicity <span class="hlt">gradient</span> of a high-redshift galaxy can vary significantly on short time-scales, associated with starburst episodes. Feedback from a starburst can destroy the gas disc, drive strong outflows and flatten a pre-existing negative metallicity <span class="hlt">gradient</span>. The time variability of a single galaxy is statistically similar to the entire simulated sample, indicating that the <span class="hlt">observed</span> metallicity <span class="hlt">gradients</span> in high-redshift galaxies reflect the instantaneous state of the galaxy rather than the accretion and growth history on cosmological time-scales. We find weak dependence of metallicity <span class="hlt">gradient</span> on stellar mass and specific star formation rate (sSFR). Low-mass galaxies and galaxies with high sSFR tend to have flat <span class="hlt">gradients</span>, likely due to the fact that feedback is more efficient in these galaxies. We argue that it is important to resolve feedback on small scales in order to produce the diverse metallicity <span class="hlt">gradients</span> <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1344439-influence-regional-nighttime-atmospheric-regimes-canopy-turbulence-gradients-closed-open-forest-mountain-valley-terrain','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1344439-influence-regional-nighttime-atmospheric-regimes-canopy-turbulence-gradients-closed-open-forest-mountain-valley-terrain"><span>Influence of regional nighttime atmospheric regimes on canopy turbulence and <span class="hlt">gradients</span> at a closed and open forest in mountain-valley terrain</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Wharton, S.; Ma, S.; Baldocchi, D. D.; ...</p> <p>2017-02-07</p> <p>Stable stratification of the nocturnal lower boundary layer inhibits convective turbulence, such that turbulent vertical transfer of ecosystem carbon dioxide (CO 2), water vapor (H 2O) and energy is driven by mechanically forced turbulence, either from frictional forces near the ground or top of a plant canopy, or from shear generated aloft. The significance of this last source of turbulence on canopy flow characteristics in a closed and open forest canopy is addressed in this paper. We present micrometeorological <span class="hlt">observations</span> of the lower boundary layer and canopy air space collected on nearly 200 nights using a combination of atmospheric lasermore » detection and ranging (lidar), eddy covariance (EC), and tower profiling instrumentation. Two AmeriFlux/Fluxnet sites in mountain-valley terrain in the Western U.S. are investigated: Wind River, a tall, dense conifer canopy, and Tonzi Ranch, a short, open oak canopy. On roughly 40% of nights lidar detected down-valley or downslope flows above the canopy at both sites. Nights with intermittent strong bursts of “top-down” forced turbulence were also <span class="hlt">observed</span> above both canopies. The strongest of these bursts increased sub-canopy turbulence and reduced canopy virtual potential temperature (θv) <span class="hlt">gradient</span> at Tonzi, but did not appear to change the flow characteristics within the dense Wind River canopy. At Tonzi we <span class="hlt">observed</span> other times when high turbulence (via friction <span class="hlt">velocity</span>, u*) was found just above the trees, yet CO2 and θv <span class="hlt">gradients</span> remained large and suggested flow decoupling. These events were triggered by regional downslope flow. Lastly, a set of turbulence parameters is evaluated for estimating canopy turbulence mixing strength. The relationship between turbulence parameters and canopy θv <span class="hlt">gradients</span> was found to be complex, although better agreement between the canopy θv <span class="hlt">gradient</span> and turbulence was found for parameters based on the standard deviation of vertical <span class="hlt">velocity</span>, or ratios of 3-D turbulence to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1344439','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1344439"><span>Influence of regional nighttime atmospheric regimes on canopy turbulence and <span class="hlt">gradients</span> at a closed and open forest in mountain-valley terrain</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wharton, S.; Ma, S.; Baldocchi, D. D.</p> <p></p> <p>Stable stratification of the nocturnal lower boundary layer inhibits convective turbulence, such that turbulent vertical transfer of ecosystem carbon dioxide (CO 2), water vapor (H 2O) and energy is driven by mechanically forced turbulence, either from frictional forces near the ground or top of a plant canopy, or from shear generated aloft. The significance of this last source of turbulence on canopy flow characteristics in a closed and open forest canopy is addressed in this paper. We present micrometeorological <span class="hlt">observations</span> of the lower boundary layer and canopy air space collected on nearly 200 nights using a combination of atmospheric lasermore » detection and ranging (lidar), eddy covariance (EC), and tower profiling instrumentation. Two AmeriFlux/Fluxnet sites in mountain-valley terrain in the Western U.S. are investigated: Wind River, a tall, dense conifer canopy, and Tonzi Ranch, a short, open oak canopy. On roughly 40% of nights lidar detected down-valley or downslope flows above the canopy at both sites. Nights with intermittent strong bursts of “top-down” forced turbulence were also <span class="hlt">observed</span> above both canopies. The strongest of these bursts increased sub-canopy turbulence and reduced canopy virtual potential temperature (θv) <span class="hlt">gradient</span> at Tonzi, but did not appear to change the flow characteristics within the dense Wind River canopy. At Tonzi we <span class="hlt">observed</span> other times when high turbulence (via friction <span class="hlt">velocity</span>, u*) was found just above the trees, yet CO2 and θv <span class="hlt">gradients</span> remained large and suggested flow decoupling. These events were triggered by regional downslope flow. Lastly, a set of turbulence parameters is evaluated for estimating canopy turbulence mixing strength. The relationship between turbulence parameters and canopy θv <span class="hlt">gradients</span> was found to be complex, although better agreement between the canopy θv <span class="hlt">gradient</span> and turbulence was found for parameters based on the standard deviation of vertical <span class="hlt">velocity</span>, or ratios of 3-D turbulence to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11969951','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11969951"><span>Planar isotropy of passive scalar turbulent mixing with a mean perpendicular <span class="hlt">gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Danaila, L; Dusek, J; Le Gal, P; Anselmet, F; Brun, C; Pumir, A</p> <p>1999-08-01</p> <p>A recently proposed evolution equation [Vaienti et al., Physica D 85, 405 (1994)] for the probability density functions (PDF's) of turbulent passive scalar increments obtained under the assumptions of fully three-dimensional homogeneity and isotropy is submitted to validation using direct numerical simulation (DNS) results of the mixing of a passive scalar with a nonzero mean <span class="hlt">gradient</span> by a homogeneous and isotropic turbulent <span class="hlt">velocity</span> field. It is shown that this approach leads to a quantitatively correct balance between the different terms of the equation, in a plane perpendicular to the mean <span class="hlt">gradient</span>, at small scales and at large Péclet number. A weaker assumption of homogeneity and isotropy restricted to the plane normal to the mean <span class="hlt">gradient</span> is then considered to derive an equation describing the evolution of the PDF's as a function of the spatial scale and the scalar increments. A very good agreement between the theory and the DNS data is obtained at all scales. As a particular case of the theory, we derive a generalized form for the well-known Yaglom equation (the isotropic relation between the second-order moments for temperature increments and the third-order <span class="hlt">velocity</span>-temperature mixed moments). This approach allows us to determine quantitatively how the integral scale properties influence the properties of mixing throughout the whole range of scales. In the simple configuration considered here, the PDF's of the scalar increments perpendicular to the mean <span class="hlt">gradient</span> can be theoretically described once the sources of inhomogeneity and anisotropy at large scales are correctly taken into account.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PApGe.174.1071P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PApGe.174.1071P"><span>On the Quality of <span class="hlt">Velocity</span> Interpolation Schemes for Marker-in-Cell Method and Staggered Grids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pusok, Adina E.; Kaus, Boris J. P.; Popov, Anton A.</p> <p>2017-03-01</p> <p>The marker-in-cell method is generally considered a flexible and robust method to model the advection of heterogenous non-diffusive properties (i.e., rock type or composition) in geodynamic problems. In this method, Lagrangian points carrying compositional information are advected with the ambient <span class="hlt">velocity</span> field on an Eulerian grid. However, <span class="hlt">velocity</span> interpolation from grid points to marker locations is often performed without considering the divergence of the <span class="hlt">velocity</span> field at the interpolated locations (i.e., non-conservative). Such interpolation schemes can induce non-physical clustering of markers when strong <span class="hlt">velocity</span> <span class="hlt">gradients</span> are present (Journal of Computational Physics 166:218-252, 2001) and this may, eventually, result in empty grid cells, a serious numerical violation of the marker-in-cell method. To remedy this at low computational costs, Jenny et al. (Journal of Computational Physics 166:218-252, 2001) and Meyer and Jenny (Proceedings in Applied Mathematics and Mechanics 4:466-467, 2004) proposed a simple, conservative <span class="hlt">velocity</span> interpolation scheme for 2-D staggered grid, while Wang et al. (Geochemistry, Geophysics, Geosystems 16(6):2015-2023, 2015) extended the formulation to 3-D finite element methods. Here, we adapt this formulation for 3-D staggered grids (correction interpolation) and we report on the quality of various <span class="hlt">velocity</span> interpolation methods for 2-D and 3-D staggered grids. We test the interpolation schemes in combination with different advection schemes on incompressible Stokes problems with strong <span class="hlt">velocity</span> <span class="hlt">gradients</span>, which are discretized using a finite difference method. Our results suggest that a conservative formulation reduces the dispersion and clustering of markers, minimizing the need of unphysical marker control in geodynamic models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26787193','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26787193"><span>Cell motility regulation on a stepped micro pillar array device (SMPAD) with a discrete stiffness <span class="hlt">gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Sujin; Hong, Juhee; Lee, Junghoon</p> <p>2016-02-28</p> <p>Our tissues consist of individual cells that respond to the elasticity of their environment, which varies between and within tissues. To better understand mechanically driven cell migration, it is necessary to manipulate the stiffness <span class="hlt">gradient</span> across a substrate. Here, we have demonstrated a new variant of the microfabricated polymeric pillar array platform that can decouple the stiffness <span class="hlt">gradient</span> from the ECM protein area. This goal is achieved via a "stepped" micro pillar array device (SMPAD) in which the contact area with the cell was kept constant while the diameter of the pillar bodies was altered to attain the proper mechanical stiffness. Using double-step SU-8 mold fabrication, the diameter of the top of every pillar was kept uniform, whereas that of the bottom was changed, to achieve the desired substrate rigidity. Fibronectin was immobilized on the pillar tops, providing a focal adhesion site for cells. C2C12, HeLa and NIH3T3 cells were cultured on the SMPAD, and the motion of the cells was <span class="hlt">observed</span> by time-lapse microscopy. Using this simple platform, which produces a purely physical stimulus, we <span class="hlt">observed</span> that various types of cell behavior are affected by the mechanical stimulus of the environment. We also demonstrated directed cell migration guided by a discrete rigidity <span class="hlt">gradient</span> by varying stiffness. Interestingly, cell <span class="hlt">velocity</span> was highest at the highest stiffness. Our approach enables the regulation of the mechanical properties of the polymeric pillar array device and eliminates the effects of the size of the contact area. This technique is a unique tool for studying cellular motion and behavior relative to various stiffness <span class="hlt">gradients</span> in the environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1361963-airborne-observations-reveal-elevational-gradient-tropical-forest-isoprene-emissions','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1361963-airborne-observations-reveal-elevational-gradient-tropical-forest-isoprene-emissions"><span>Airborne <span class="hlt">observations</span> reveal elevational <span class="hlt">gradient</span> in tropical forest isoprene emissions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Gu, Dasa; Guenther, Alex B.; Shilling, John E.; ...</p> <p>2017-05-23</p> <p>Terrestrial vegetation emits vast quantities of volatile organic compounds (VOCs) to he atmosphere1-3, which influence oxidants and aerosols leading to complex feedbacks on air quality and climate4-6. Isoprene dominates global non-methane VOC emissions with tropical regions contributing ~80% of global isoprene emissions2. Isoprene emission rates vary over several orders of magnitude for different plant species, and characterizing this immense biological chemodiversity is a challenge for estimating isoprene emission from tropical forests. Here we present the isoprene emission estimates from aircraft direct eddy covariance measurements over the pristine Amazon forest. We report isoprene emission rates that are 3 times higher thanmore » satellite top-down estimates and 35% higher than model predictions based on satellite land cover and vegetation specific emission factors (EFs). The results reveal strong correlations between <span class="hlt">observed</span> isoprene emission rates and terrain elevations which are confirmed by similar correlations between satellite-derived isoprene emissions and terrain elevations. We propose that the elevational <span class="hlt">gradient</span> in the Amazonian forest isoprene emission capacity is determined by plant species distributions and can explain a substantial degree of isoprene emission variability in tropical forests. Finally, we apply this approach over the central Amazon and use a model to demonstrate the impacts on regional air quality.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDQ28001L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDQ28001L"><span>Universality of the logarithmic <span class="hlt">velocity</span> profile restored</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luchini, Paolo</p> <p>2017-11-01</p> <p>The logarithmic <span class="hlt">velocity</span> profile of wall-bounded turbulent flow, despite its widespread adoption in research and in teaching, exhibits discrepancies with both experiments and numerical simulations that have been repeatedly <span class="hlt">observed</span> in the literature; serious doubts ensued about its precise form and universality, leading to the formulation of alternate theories and hindering ongoing experimental efforts to measure von Kármán's constant. By comparing different geometries of pipe, plane-channel and plane-Couette flow, here we show that such discrepancies can be physically interpreted, and analytically accounted for, through an equally universal higher-order correction caused by the pressure <span class="hlt">gradient</span>. Inclusion of this term produces a tenfold increase in the adherence of the predicted profile to existing experiments and numerical simulations in all three geometries. Universality of the logarithmic law then emerges beyond doubt and a satisfactorily simple formulation is established. Among the consequences of this formulation is a strongly increased confidence that the Reynolds number of present-day direct numerical simulations is actually high enough to uncover asymptotic behaviour, but research efforts are still needed in order to increase their accuracy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040141463&hterms=left&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dleft','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040141463&hterms=left&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dleft"><span>Doppler echo evaluation of pulmonary venous-left atrial pressure <span class="hlt">gradients</span>: human and numerical model studies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Firstenberg, M. S.; Greenberg, N. L.; Smedira, N. G.; Prior, D. L.; Scalia, G. M.; Thomas, J. D.; Garcia, M. J.</p> <p>2000-01-01</p> <p>The simplified Bernoulli equation relates fluid convective energy derived from flow <span class="hlt">velocities</span> to a pressure <span class="hlt">gradient</span> and is commonly used in clinical echocardiography to determine pressure differences across stenotic orifices. Its application to pulmonary venous flow has not been described in humans. Twelve patients undergoing cardiac surgery had simultaneous high-fidelity pulmonary venous and left atrial pressure measurements and pulmonary venous pulsed Doppler echocardiography performed. Convective <span class="hlt">gradients</span> for the systolic (S), diastolic (D), and atrial reversal (AR) phases of pulmonary venous flow were determined using the simplified Bernoulli equation and correlated with measured actual pressure differences. A linear relationship was <span class="hlt">observed</span> between the convective (y) and actual (x) pressure differences for the S (y = 0.23x + 0.0074, r = 0.82) and D (y = 0.22x + 0.092, r = 0.81) waves, but not for the AR wave (y = 0. 030x + 0.13, r = 0.10). Numerical modeling resulted in similar slopes for the S (y = 0.200x - 0.127, r = 0.97), D (y = 0.247x - 0. 354, r = 0.99), and AR (y = 0.087x - 0.083, r = 0.96) waves. Consistent with numerical modeling, the convective term strongly correlates with but significantly underestimates actual <span class="hlt">gradient</span> because of large inertial forces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10924058','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10924058"><span>Doppler echo evaluation of pulmonary venous-left atrial pressure <span class="hlt">gradients</span>: human and numerical model studies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Firstenberg, M S; Greenberg, N L; Smedira, N G; Prior, D L; Scalia, G M; Thomas, J D; Garcia, M J</p> <p>2000-08-01</p> <p>The simplified Bernoulli equation relates fluid convective energy derived from flow <span class="hlt">velocities</span> to a pressure <span class="hlt">gradient</span> and is commonly used in clinical echocardiography to determine pressure differences across stenotic orifices. Its application to pulmonary venous flow has not been described in humans. Twelve patients undergoing cardiac surgery had simultaneous high-fidelity pulmonary venous and left atrial pressure measurements and pulmonary venous pulsed Doppler echocardiography performed. Convective <span class="hlt">gradients</span> for the systolic (S), diastolic (D), and atrial reversal (AR) phases of pulmonary venous flow were determined using the simplified Bernoulli equation and correlated with measured actual pressure differences. A linear relationship was <span class="hlt">observed</span> between the convective (y) and actual (x) pressure differences for the S (y = 0.23x + 0.0074, r = 0.82) and D (y = 0.22x + 0.092, r = 0.81) waves, but not for the AR wave (y = 0. 030x + 0.13, r = 0.10). Numerical modeling resulted in similar slopes for the S (y = 0.200x - 0.127, r = 0.97), D (y = 0.247x - 0. 354, r = 0.99), and AR (y = 0.087x - 0.083, r = 0.96) waves. Consistent with numerical modeling, the convective term strongly correlates with but significantly underestimates actual <span class="hlt">gradient</span> because of large inertial forces.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AIPC.1333..301M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AIPC.1333..301M"><span>Breakdown and Limit of Continuum Diffusion <span class="hlt">Velocity</span> for Binary Gas Mixtures from Direct Simulation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Martin, Robert Scott; Najmabadi, Farrokh</p> <p>2011-05-01</p> <p>This work investigates the breakdown of the continuum relations for diffusion <span class="hlt">velocity</span> in inert binary gas mixtures. Values of the relative diffusion <span class="hlt">velocities</span> for components of a gas mixture may be calculated using of Chapman-Enskog theory and occur not only due to concentration <span class="hlt">gradients</span>, but also pressure and temperature <span class="hlt">gradients</span> in the flow as described by Hirschfelder. Because Chapman-Enskog theory employs a linear perturbation around equilibrium, it is expected to break down when the <span class="hlt">velocity</span> distribution deviates significantly from equilibrium. This breakdown of the overall flow has long been an area of interest in rarefied gas dynamics. By comparing the continuum values to results from Bird's DS2V Monte Carlo code, we propose a new limit on the continuum approach specific to binary gases. To remove the confounding influence of an inconsistent molecular model, we also present the application of the variable hard sphere (VSS) model used in DS2V to the continuum diffusion <span class="hlt">velocity</span> calculation. Fitting sample asymptotic curves to the breakdown, a limit, Vmax, that is a fraction of an analytically derived limit resulting from the kinetic temperature of the mixture is proposed. With an expected deviation of only 2% between the physical values and continuum calculations within ±Vmax/4, we suggest this as a conservative estimate on the range of applicability for the continuum theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70017889','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70017889"><span>Application of Conjugate <span class="hlt">Gradient</span> methods to tidal simulation</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Barragy, E.; Carey, G.F.; Walters, R.A.</p> <p>1993-01-01</p> <p>A harmonic decomposition technique is applied to the shallow water equations to yield a complex, nonsymmetric, nonlinear, Helmholtz type problem for the sea surface and an accompanying complex, nonlinear diagonal problem for the <span class="hlt">velocities</span>. The equation for the sea surface is linearized using successive approximation and then discretized with linear, triangular finite elements. The study focuses on applying iterative methods to solve the resulting complex linear systems. The comparative evaluation includes both standard iterative methods for the real subsystems and complex versions of the well known Bi-Conjugate <span class="hlt">Gradient</span> and Bi-Conjugate <span class="hlt">Gradient</span> Squared methods. Several Incomplete LU type preconditioners are discussed, and the effects of node ordering, rejection strategy, domain geometry and Coriolis parameter (affecting asymmetry) are investigated. Implementation details for the complex case are discussed. Performance studies are presented and comparisons made with a frontal solver. ?? 1993.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999GeoJI.138..871R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999GeoJI.138..871R"><span>Lithospheric structure of the Arabian Shield and Platform from complete regional waveform modelling and surface wave group <span class="hlt">velocities</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodgers, Arthur J.; Walter, William R.; Mellors, Robert J.; Al-Amri, Abdullah M. S.; Zhang, Yu-Shen</p> <p>1999-09-01</p> <p>Regional seismic waveforms reveal significant differences in the structure of the Arabian Shield and the Arabian Platform. We estimate lithospheric <span class="hlt">velocity</span> structure by modelling regional waveforms recorded by the 1995-1997 Saudi Arabian Temporary Broadband Deployment using a grid search scheme. We employ a new method whereby we narrow the waveform modelling grid search by first fitting the fundamental mode Love and Rayleigh wave group <span class="hlt">velocities</span>. The group <span class="hlt">velocities</span> constrain the average crustal thickness and <span class="hlt">velocities</span> as well as the crustal <span class="hlt">velocity</span> <span class="hlt">gradients</span>. Because the group <span class="hlt">velocity</span> fitting is computationally much faster than the synthetic seismogram calculation this method allows us to determine good average starting models quickly. Waveform fits of the Pn and Sn body wave arrivals constrain the mantle <span class="hlt">velocities</span>. The resulting lithospheric structures indicate that the Arabian Platform has an average crustal thickness of 40 km, with relatively low crustal <span class="hlt">velocities</span> (average crustal P- and S-wave <span class="hlt">velocities</span> of 6.07 and 3.50 km s^-1 , respectively) without a strong <span class="hlt">velocity</span> <span class="hlt">gradient</span>. The Moho is shallower (36 km) and crustal <span class="hlt">velocities</span> are 6 per cent higher (with a <span class="hlt">velocity</span> increase with depth) for the Arabian Shield. Fast crustal <span class="hlt">velocities</span> of the Arabian Shield result from a predominantly mafic composition in the lower crust. Lower <span class="hlt">velocities</span> in the Arabian Platform crust indicate a bulk felsic composition, consistent with orogenesis of this former active margin. P- and S-wave <span class="hlt">velocities</span> immediately below the Moho are slower in the Arabian Shield than in the Arabian Platform (7.9 and 4.30 km s^-1 , and 8.10 and 4.55 km s^-1 , respectively). This indicates that the Poisson's ratios for the uppermost mantle of the Arabian Shield and Platform are 0.29 and 0.27, respectively. The lower mantle <span class="hlt">velocities</span> and higher Poisson's ratio beneath the Arabian Shield probably arise from a partially molten mantle associated with Red Sea spreading and continental</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12069920','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12069920"><span>Influence of natural temperature <span class="hlt">gradients</span> on measurements of xylem sap flow with thermal dissipation probes. 1. Field <span class="hlt">observations</span> and possible remedies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Do, F; Rocheteau, A</p> <p>2002-06-01</p> <p>The thermal dissipation method is simple and widely used for measuring sap flow in large stems. As with several other thermal methods, natural temperature <span class="hlt">gradients</span> are assumed to be negligible in the sapwood being measured. We studied the magnitude and variability of natural temperature <span class="hlt">gradients</span> in sapwood of Acacia trees growing in the Sahelian zone of Senegal, analyzed their effects on sap flow measurements, and investigated possible solutions. A new measurement approach employing cyclic heating (45 minutes of heating and 15 minutes of cooling; 45/15) was also tested. Three-day measurement sequences that included 1 day without heating, a second day with continuous heating and a third day with cyclic heating were recorded during a 6.5-month period using probes installed at three azimuths in a tree trunk. Natural temperature <span class="hlt">gradients</span> between the two probes of the sensor unit, spaced 8 to 10 cm vertically, were rarely negligible (i.e., < 0.2 degrees C): they were positive during the night and negative during the day, with an amplitude ranging from 0.3 to 3.5 degrees C depending on trunk azimuth, day and season. These temperature <span class="hlt">gradients</span> had a direct influence on the signal from the continuously heated sensors, inducing fluctuations in the nighttime reference signal. The resulting errors in sap flow estimates can be greater than 100%. Correction protocols have been proposed in previous studies, but they were unsuitable because of the high spatial and temporal variability of the natural temperature <span class="hlt">gradients</span>. We found that a measurement signal derived from a noncontinuous heating system could be an attractive solution because it appears to be independent of natural temperature <span class="hlt">gradients</span>. The magnitude and variability of temperature <span class="hlt">gradients</span> that we <span class="hlt">observed</span> were likely exacerbated by the combination of open stand, high solar radiation and low sap flow rate. However, for all applications of the thermal dissipation method, it is wise to check regularly for natural</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMGC23L1258G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMGC23L1258G"><span>Spatio-temporal variability of vertical <span class="hlt">gradients</span> of major meteorological <span class="hlt">observations</span> around the Tibetan Plateau</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, X.; Wang, L.; Tian, L.</p> <p>2015-12-01</p> <p>The near-surface air temperature lapse rate (TLR), wind speed <span class="hlt">gradient</span> (WSG), and precipitation <span class="hlt">gradient</span> (PG) provide crucial parameters used in models of mountain climate and hydrology. The complex mountain terrain and vast area of the Tibetan Plateau (TP) make such factors particularly important. With daily data from 161 meteorological stations over the past 43 years (1970-2012), we analyse the spatio-temporal variations of TLRs, WSGs, and PGs over and around TP, derived using linear regression methods and dividing the study area into zones based on spatial variations. Results of this study include: (1) The <span class="hlt">observed</span> TLR varies from -0.46 to -0.73 ∘C (100 m) -1, with averaged TLRs of -0.60,-0.62, and -0.59 ∘C (100 m) -1 for Tmax, Tmin,and Tmean , respectively. The averaged TLR is slightly less than the global mean of -0.65 ∘C (100 m) -1 . The spatial variability of TLR relates to climate conditions, wherein the TLR increases in dry conditions and in cold months (October-April), while it lessens in humid regions and during warm months (May-September). (2) The estimated annual WSG ranges from 0.07 to 0.17m s -1 (100 m) -1. Monthly WSGs show a marked seasonal shift, in which higher WSGs can be explained by the high intensity of prevailing wind. (3) Positive summer PGs vary from 12.08 in the central TP to 26.14 mm (100 m) -1 in northeastern Qinghai and the southern TP, but a reverse <span class="hlt">gradient</span> prevails in Yunnan and parts of Sichuan Province. (4) The regional warming over TP is more evident in winter, and Tmin demonstrated the most prominent warming compared with Tmax and Tmean. Environments at high elevations experience more rapid changes in temperatures (Tmax, Tmin,and Tmean) than those at low elevations, which is especially true in winter and for Tmin. Furthermore, inter-annual variation of TLRs is linked to elevation-dependent warming.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011APS..MARD32001D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011APS..MARD32001D"><span>Spatial <span class="hlt">gradient</span> tuning in metamaterials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Driscoll, Tom; Goldflam, Michael; Jokerst, Nan; Basov, Dimitri; Smith, David</p> <p>2011-03-01</p> <p><span class="hlt">Gradient</span> Index (GRIN) metamaterials have been used to create devices inspired by, but often surpassing the potential of, conventional GRIN optics. The unit-cell nature of metamaterials presents the opportunity to exert much greater control over spatial <span class="hlt">gradients</span> than is possible in natural materials. This is true not only during the design phase but also offers the potential for real-time reconfiguration of the metamaterial <span class="hlt">gradient</span>. This ability fits nicely into the picture of transformation-optics, in which spatial <span class="hlt">gradients</span> can enable an impressive suite of innovative devices. We discuss methods to exert control over metamaterial response, focusing on our recent demonstrations using Vanadium Dioxide. We give special attention to role of memristance and mem-capacitance <span class="hlt">observed</span> in Vanadium Dioxide, which simplify the demands of stimuli and addressing, as well as intersecting metamaterials with the field of memory-materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDR41011M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDR41011M"><span>Decay and Spatial Diffusion of Turbulent Kinetic Energy In The Presence of a Linear Kinetic Energy <span class="hlt">Gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meneveau, Charles</p> <p>2015-11-01</p> <p>A topic that elicited the interest of John Lumley is pressure transport in turbulence. In 1978 (JL, in Advances in Applied Mechanics, pages 123-176) he showed that pressure transport likely acts in the opposite direction to the spatial flux of kinetic energy due to triple <span class="hlt">velocity</span> correlations. Here we examine a flow in which the interplay of turbulent decay and spatial transport is particularly relevant. Specifically, using a specially designed active grid and screens placed in the Corrsin wind tunnel, such a flow is realized. Data are acquired using X-wire thermal anemometry at different spanwise and downstream locations. In order to resolve the dissipation rate accurately, measurements are also acquired using the NSTAP probe developed and manufactured by Princeton researchers and kindly provided to us (M. Hultmark, Y. Fan, L. Smits). The results show power-law decay with downstream distance, with a decay exponent that becomes larger in the high kinetic energy side of the flow. Measurements of the dissipation enable us to obtain the spanwise <span class="hlt">gradient</span> of the spatial flux. One possible explanation for the <span class="hlt">observations</span> is upgrading transport of kinetic energy due to pressure-<span class="hlt">velocity</span> correlations, although its magnitude required to close the budget appears very large. Absence of simultaneous pressure <span class="hlt">velocity</span> measurement preclude us to fully elucidate the <span class="hlt">observed</span> trends. In collaboration with Adrien Thormann, Johns Hopkins University. Financial support: National Science Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28750164','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28750164"><span>Fast Transport of Water Droplets over a Thermo-Switchable Surface Using Rewritable Wettability <span class="hlt">Gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Banuprasad, Theneyur Narayanaswamy; Vinay, Thamarasseril Vijayan; Subash, Cherumannil Karumuthil; Varghese, Soney; George, Sajan D; Varanakkottu, Subramanyan Namboodiri</p> <p>2017-08-23</p> <p>In spite of the reported temperature dependent tunability in wettability of poly(N-isopropylacrylamide) (PNIPAAm) surfaces for below and above lower critical solution temperature (32 °C), the transport of water droplets is inhibited by the large contact angle hysteresis. Herein, for the first time, we report on-demand, fast, and reconfigurable droplet manipulation over a PNIPAAm grafted structured polymer surface using temperature-induced wettability <span class="hlt">gradient</span>. Our study reveals that the PNIPAAm grafted on intrinsically superhydrophobic surfaces exhibit hydrophilic nature with high contact angle hysteresis below 30 °C and superhydrophobic nature with ultralow contact angle hysteresis above 36 °C. The transition region between 30 and 36 °C is characterized by a large change in water contact angle (∼100°) with a concomitant change in contact angle hysteresis. By utilizing this "transport zone" wherein driving forces overcome the frictional forces, we demonstrate macroscopic transport of water drops with a maximum transport <span class="hlt">velocity</span> of approximately 40 cm/s. The theoretical calculations on the force measurements concur with dominating behavior of driving forces across the transport zone. The tunability in transport <span class="hlt">velocity</span> by varying the temperature <span class="hlt">gradient</span> along the surface or the inclination angle of the surface (maximum angle of 15° with a reduced <span class="hlt">velocity</span> 0.4 mm/s) is also elucidated. In addition, as a practical application, coalescence of water droplets is demonstrated by using the temperature controlled wettability <span class="hlt">gradient</span>. The presented results are expected to provide new insights on the design and fabrication of smart multifunctional surfaces for applications such as biochemical analysis, self-cleaning, and microfluidics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PApGe.173.2387G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PApGe.173.2387G"><span>Present-Day 3D <span class="hlt">Velocity</span> Field of Eastern North America Based on Continuous GPS <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goudarzi, Mohammad Ali; Cocard, Marc; Santerre, Rock</p> <p>2016-07-01</p> <p>The Saint Lawrence River valley in eastern Canada was studied using <span class="hlt">observations</span> of continuously operating GPS (CGPS) stations. The area is one of the most seismically active regions in eastern North America characterized by many earthquakes, which is also subject to an ongoing glacial isostatic adjustment. We present the current three-dimensional <span class="hlt">velocity</span> field of eastern North America obtained from more than 14 years (9 years on average) of data at 112 CGPS stations. Bernese GNSS and GITSA software were used for CGPS data processing and position time series analysis, respectively. The results show the counterclockwise rotation of the North American plate in the No-Net-Rotation model with the average of 16.8 ± 0.7 mm/year constrained to ITRF 2008. We also present an ongoing uplift model for the study region based on the present-day CGPS <span class="hlt">observations</span>. The model shows uplift all over eastern Canada with the maximum rate of 13.7 ± 1.2 mm/year and subsidence to the south mainly over northern USA with a typical rate of -1 to -2 mm/year and the minimum value of -2.7 ± 1.4 mm/year. We compared our model with the rate of radial displacements from the ICE-5G model. Both models agree within 0.02 mm/year at the best stations; however, our model shows a systematic spatial tilt compared to ICE-5G. The misfits between two models amount to the maximum relative subsidence of -6.1 ± 1.1 mm/year to the east and maximum relative uplift of 5.9 ± 2.7 mm/year to the west. The intraplate horizontal <span class="hlt">velocities</span> are radially outward from the centers of maximum uplift and are inward to the centers of maximum subsidence with the typical <span class="hlt">velocity</span> of 1-1.6 ± 0.4 mm/year that is in agreement with the ICE-5G model to the first order.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999A%26A...341...86B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999A%26A...341...86B"><span>The local stellar <span class="hlt">velocity</span> distribution of the Galaxy. Galactic structure and potential</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bienaymé, O.</p> <p>1999-01-01</p> <p>The <span class="hlt">velocity</span> distribution of neighbouring stars is deduced from the Hipparcos proper motions. We have used a classical Schwarzschild decomposition and also developed a dynamical model for quasi-exponential stellar discs. This model is a 3-D derivation of Shu's model in the framework of Stäckel potentials with three integrals of motion. We determine the solar motion relative to the local standard of rest (LSR) (U_sun=9.7+/-0.3kms , V_sun=5.2+/-1.0kms and W_sun=6.7+/-0.2kms ), the density and kinematic radial <span class="hlt">gradients</span>, as well as the local slope of the <span class="hlt">velocity</span> curve. We find out that the scale density length of the Galaxy is 1.8+/-0.2kpc . We measure a large kinematic scale length for blue (young) stars, R_{sigma_r }=17+/-4kpc , while for red stars (predominantly old) we find R_{sigma_r }=9.7+/-0.8kpc (or R_{sigma_r (2}=4.8+/-0.4kpc ) ). From the stellar disc dynamical model, we determine explicitly the link between the tangential-vertical <span class="hlt">velocity</span> (v_theta , v_z) coupling and the local shape of the potential. Using a restricted sample of 3-D <span class="hlt">velocity</span> data, we measure z_o, the focus of the spheroidal coordinate system defining the best fitted Stäckel potential. The parameter z_o is related to the tilt of the <span class="hlt">velocity</span> ellipsoid and more fundamentally to the mass <span class="hlt">gradient</span> in the galactic disc. This parameter is found to be 5.7+/-1.4kpc . This implies that the galactic potential is not extremely flat and that the dark matter component is not confined in the galactic plane. Based on data from the Hipparcos astrometry satellite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1438438-nonlinear-verification-linear-critical-gradient-model-energetic-particle-transport-alfven-eigenmodes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1438438-nonlinear-verification-linear-critical-gradient-model-energetic-particle-transport-alfven-eigenmodes"><span>Nonlinear verification of a linear critical <span class="hlt">gradient</span> model for energetic particle transport by Alfven eigenmodes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bass, Eric M.; Waltz, R. E.</p> <p></p> <p>Here, a “stiff transport” critical <span class="hlt">gradient</span> model of energetic particle (EP) transport by EPdriven Alfven eigenmodes (AEs) is verified against local nonlinear gyrokinetic simulations of a well-studied beam-heated DIII-D discharge 146102. A greatly simplifying linear “recipe” for the limiting EP-density <span class="hlt">gradient</span> (critical <span class="hlt">gradient</span>) is considered here. In this recipe, the critical <span class="hlt">gradient</span> occurs when the AE linear growth rate, driven mainly by the EP <span class="hlt">gradient</span>, exceeds the ion temperature <span class="hlt">gradient</span> (ITG) or trapped electron mode (TEM) growth rate, driven by the thermal plasma <span class="hlt">gradient</span>, at the same toroidal mode number (n) as the AE peak growth, well below the ITG/TEMmore » peak n. This linear recipe for the critical <span class="hlt">gradient</span> is validated against the critical <span class="hlt">gradient</span> determined from far more expensive local nonlinear simulations in the gyrokinetic code GYRO, as identified by the point of transport runaway when all driving <span class="hlt">gradients</span> are held fixed. The reduced linear model is extended to include the stabilization from equilibrium E×B <span class="hlt">velocity</span> shear. The nonlinear verification unambiguously endorses one of two alternative recipes proposed in Ref. 1: the EP-driven AE growth rate should be determined with rather than without added thermal plasma drive.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1438438-nonlinear-verification-linear-critical-gradient-model-energetic-particle-transport-alfven-eigenmodes','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1438438-nonlinear-verification-linear-critical-gradient-model-energetic-particle-transport-alfven-eigenmodes"><span>Nonlinear verification of a linear critical <span class="hlt">gradient</span> model for energetic particle transport by Alfven eigenmodes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Bass, Eric M.; Waltz, R. E.</p> <p>2017-12-08</p> <p>Here, a “stiff transport” critical <span class="hlt">gradient</span> model of energetic particle (EP) transport by EPdriven Alfven eigenmodes (AEs) is verified against local nonlinear gyrokinetic simulations of a well-studied beam-heated DIII-D discharge 146102. A greatly simplifying linear “recipe” for the limiting EP-density <span class="hlt">gradient</span> (critical <span class="hlt">gradient</span>) is considered here. In this recipe, the critical <span class="hlt">gradient</span> occurs when the AE linear growth rate, driven mainly by the EP <span class="hlt">gradient</span>, exceeds the ion temperature <span class="hlt">gradient</span> (ITG) or trapped electron mode (TEM) growth rate, driven by the thermal plasma <span class="hlt">gradient</span>, at the same toroidal mode number (n) as the AE peak growth, well below the ITG/TEMmore » peak n. This linear recipe for the critical <span class="hlt">gradient</span> is validated against the critical <span class="hlt">gradient</span> determined from far more expensive local nonlinear simulations in the gyrokinetic code GYRO, as identified by the point of transport runaway when all driving <span class="hlt">gradients</span> are held fixed. The reduced linear model is extended to include the stabilization from equilibrium E×B <span class="hlt">velocity</span> shear. The nonlinear verification unambiguously endorses one of two alternative recipes proposed in Ref. 1: the EP-driven AE growth rate should be determined with rather than without added thermal plasma drive.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JAP...123i1702L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JAP...123i1702L"><span>Inherent losses induced absorptive acoustic rainbow trapping with a <span class="hlt">gradient</span> metasurface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Tuo; Liang, Shanjun; Chen, Fei; Zhu, Jie</p> <p>2018-03-01</p> <p>Acoustic rainbow trapping represents the phenomenon of strong acoustic dispersion similar to the optical "trapped rainbow," which allows spatial-spectral modulation and broadband trapping of sound. It can be realized with metamaterials that provide the required strong dispersion absent in natural materials. However, as the group <span class="hlt">velocity</span> cannot be reduced to exactly zero before the forward mode being coupled to the backward mode, such trapping is temporary and the local sound oscillation ultimately radiates backward. Here, we propose a <span class="hlt">gradient</span> metasurface, a rigid surface structured with <span class="hlt">gradient</span> perforation along the wave propagation direction, in which the inherent thermal and viscous losses inside the holes are considered. We show that the gradually diminished group <span class="hlt">velocity</span> of the structure-induced surface acoustic waves (SSAWs) supported by the metasurface becomes anomalous at the trapping position, induced by the existence of the inherent losses, which implies that the system's absorption reaches its maximum. Together with the progressively increased attenuation of the SSAWs along the <span class="hlt">gradient</span> direction, reflectionless spatial-spectral modulation and sound enhancement are achieved in simulation. Such phenomenon, which we call as absorptive trapped rainbow, results from the balanced interplay among the local resonance inside individual holes, the mutual coupling of adjacent unit cells, and the inherent losses due to thermal conductivity and viscosity. This study deepens the understanding of the SSAWs propagation at a lossy metasurface and may contribute to the practical design of acoustic devices for high performance sensing and filtering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFDA25002K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFDA25002K"><span>Plasma Streamwise Vortex Generators in an Adverse Pressure <span class="hlt">Gradient</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kelley, Christopher; Corke, Thomas; Thomas, Flint</p> <p>2013-11-01</p> <p>A wind tunnel experiment was conducted to compare plasma streamwise vortex generators (PSVGs) and passive vortex generators (VGs). These devices were installed on a wing section by which the angle of attack could be used to vary the streamwise pressure <span class="hlt">gradient</span>. The experiment was performed for freestream Mach numbers 0.1-0.2. Three-dimensional <span class="hlt">velocity</span> components were measured using a 5-hole Pitot probe in the boundary layer. These measurements were used to quantify the production of streamwise vorticity and the magnitude of the reorientation term from the vorticity transport equation. The effect of Mach number, pressure <span class="hlt">gradient</span>, operating voltage, and electrode length was then investigated for the PSVGs. The results indicate that the PSVGs could easily outperform the passive VGs and provide a suitable alternative for flow control.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110013647','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110013647"><span>Crosswind Shear <span class="hlt">Gradient</span> Affect on Wake Vortices</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Proctor, Fred H.; Ahmad, Nashat N.</p> <p>2011-01-01</p> <p>Parametric simulations with a Large Eddy Simulation (LES) model are used to explore the influence of crosswind shear on aircraft wake vortices. Previous studies based on field measurements, laboratory experiments, as well as LES, have shown that the vertical <span class="hlt">gradient</span> of crosswind shear, i.e. the second vertical derivative of the environmental crosswind, can influence wake vortex transport. The presence of nonlinear vertical shear of the crosswind <span class="hlt">velocity</span> can reduce the descent rate, causing a wake vortex pair to tilt and change in its lateral separation. The LES parametric studies confirm that the vertical <span class="hlt">gradient</span> of crosswind shear does influence vortex trajectories. The parametric results also show that vortex decay from the effects of shear are complex since the crosswind shear, along with the vertical <span class="hlt">gradient</span> of crosswind shear, can affect whether the lateral separation between wake vortices is increased or decreased. If the separation is decreased, the vortex linking time is decreased, and a more rapid decay of wake vortex circulation occurs. If the separation is increased, the time to link is increased, and at least one of the vortices of the vortex pair may have a longer life time than in the case without shear. In some cases, the wake vortices may never link.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15376943','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15376943"><span><span class="hlt">Gradient</span>-based multiresolution image fusion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Petrović, Valdimir S; Xydeas, Costas S</p> <p>2004-02-01</p> <p>A novel approach to multiresolution signal-level image fusion is presented for accurately transferring visual information from any number of input image signals, into a single fused image without loss of information or the introduction of distortion. The proposed system uses a "fuse-then-decompose" technique realized through a novel, fusion/decomposition system architecture. In particular, information fusion is performed on a multiresolution <span class="hlt">gradient</span> map representation domain of image signal information. At each resolution, input images are represented as <span class="hlt">gradient</span> maps and combined to produce new, fused <span class="hlt">gradient</span> maps. Fused <span class="hlt">gradient</span> map signals are processed, using <span class="hlt">gradient</span> filters derived from high-pass quadrature mirror filters to yield a fused multiresolution pyramid representation. The fused output image is obtained by applying, on the fused pyramid, a reconstruction process that is analogous to that of conventional discrete wavelet transform. This new <span class="hlt">gradient</span> fusion significantly reduces the amount of distortion artefacts and the loss of contrast information usually <span class="hlt">observed</span> in fused images obtained from conventional multiresolution fusion schemes. This is because fusion in the <span class="hlt">gradient</span> map domain significantly improves the reliability of the feature selection and information fusion processes. Fusion performance is evaluated through informal visual inspection and subjective psychometric preference tests, as well as objective fusion performance measurements. Results clearly demonstrate the superiority of this new approach when compared to conventional fusion systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22089794-keck-nirspec-radial-velocity-observations-late-dwarfs','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22089794-keck-nirspec-radial-velocity-observations-late-dwarfs"><span>KECK NIRSPEC RADIAL <span class="hlt">VELOCITY</span> <span class="hlt">OBSERVATIONS</span> OF LATE-M DWARFS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Tanner, Angelle; White, Russel; Bailey, John</p> <p>2012-11-15</p> <p>We present the results of an infrared spectroscopic survey of 23 late-M dwarfs with the NIRSPEC echelle spectrometer on the Keck II telescope. Using telluric lines for wavelength calibration, we are able to achieve measurement precisions of down to 45 m s{sup -1} for our late-M dwarfs over a one- to four-year long baseline. Our sample contains two stars with radial <span class="hlt">velocity</span> (RV) variations of >1000 m s{sup -1}. While we require more measurements to determine whether these RV variations are due to unseen planetary or stellar companions or are the result of starspots known to plague the surface ofmore » M dwarfs, we can place upper limits of <40 M{sub J} sin i on the masses of any companions around those two M dwarfs with RV variations of <160 m s{sup -1} at orbital periods of 10-100 days. We have also measured the rotational <span class="hlt">velocities</span> for all the stars in our late-M dwarf sample and offer our multi-order, high-resolution spectra over 2.0-2.4 {mu}m to the atmospheric modeling community to better understand the atmospheres of late-M dwarfs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930010063','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930010063"><span><span class="hlt">Velocity</span> distribution of fragments of catastrophic impacts</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Takagi, Yasuhiko; Kato, Manabu; Mizutani, Hitoshi</p> <p>1992-01-01</p> <p>Three dimensional <span class="hlt">velocities</span> of fragments produced by laboratory impact experiments were measured for basalts and pyrophyllites. The <span class="hlt">velocity</span> distribution of fragments obtained shows that the <span class="hlt">velocity</span> range of the major fragments is rather narrow, at most within a factor of 3 and that no clear dependence of <span class="hlt">velocity</span> on the fragment mass is <span class="hlt">observed</span>. The NonDimensional Impact Stress (NDIS) defined by Mizutani et al. (1990) is found to be an appropriate scaling parameter to describe the overall fragment <span class="hlt">velocity</span> as well as the antipodal <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19730051463&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D60%26Ntt%3Dlazarus','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19730051463&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D60%26Ntt%3Dlazarus"><span><span class="hlt">Observation</span> and analysis of abrupt changes in the interplanetary plasma <span class="hlt">velocity</span> and magnetic field.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Martin, R. N.; Belcher, J. W.; Lazarus, A. J.</p> <p>1973-01-01</p> <p>This paper presents a limited study of the physical nature of abrupt changes in the interplanetary plasma <span class="hlt">velocity</span> and magnetic field based on 19 day's data from the Pioneer 6 spacecraft. The period was chosen to include a high-<span class="hlt">velocity</span> solar wind stream and low-<span class="hlt">velocity</span> wind. Abrupt events were accepted for study if the sum of the energy density in the magnetic field and <span class="hlt">velocity</span> changes was above a specified minimum. A statistical analysis of the events in the high-<span class="hlt">velocity</span> solar wind stream shows that Alfvenic changes predominate. This conclusion is independent of whether steady state requirements are imposed on conditions before and after the event. Alfvenic changes do not dominate in the lower-speed wind. This study extends the plasma field evidence for outwardly propagating Alfvenic changes to time scales as small as 1 min (scale lengths on the order of 20,000 km).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AAS...21812712Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AAS...21812712Y"><span>High <span class="hlt">Velocity</span> Precessing Jet from the Water Fountain IRAS 18286-0959 Revealed by VLBA <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yung, Bosco; Nakashima, J.; Imai, H.; Deguchi, S.; Diamond, P. J.; Kwok, S.</p> <p>2011-05-01</p> <p>We report the multi-epoch VLBA <span class="hlt">observations</span> of 22.2GHz water maser emission associated with the "water fountain" star IRAS 18286-0959. The detected maser emission are distributed in the <span class="hlt">velocity</span> range from -50km/s to 150km/s. The spatial distribution of over 70% of the identified maser features is found to be highly collimated along a spiral jet (namely, jet 1) extended from southeast to northwest direction, and the rest of the features appear to trace another spiral jet (jet 2) with a different orientation. The two jets form a "double-helix" pattern which lies across 200 milliarcseconds (mas). The maser features are reasonably fit by a model consisting of two precessing jets. The <span class="hlt">velocities</span> of jet 1 and jet 2 are derived to be 138km/s and 99km/s, respectively. The precession period of jet 1 is about 56 years, and for jet 2 it is about 73 years. We propose that the appearance of two jets <span class="hlt">observed</span> are the result of a single driving source with a significant proper motion. This research was supported by grants from the Research Grants Council of the Hong Kong Special Administrative Region, China, the Seed Funding Programme for Basic Research of the University of Hong Kong, Grant-in-Aid for Young Scientists from the Ministry 9 of Education, Culture, Sports, Science, and Technology, and Grant-in-Aid for Scientific Research from Japan Society for Promotion Science.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDL22008C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDL22008C"><span>Assessment of fluctuating pressure <span class="hlt">gradient</span> using acceleration spectra in near wall flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cadel, Daniel; Lowe, K. Todd</p> <p>2015-11-01</p> <p>Separation of contributions to the fluctuating acceleration from pressure <span class="hlt">gradient</span> fluctuations and viscous shear fluctuations in the frequency domain is examined in a turbulent boundary layer. Past work leveraging turbulent accelerations for pressure <span class="hlt">gradient</span> measurements has neglected the viscous shear term from the momentum equation--an invalid assumption in the case of near wall flows. The present study seeks to account for the influence of the viscous shear term and spectrally reject its contribution, which is thought to be concentrated at higher frequencies. Spectra of <span class="hlt">velocity</span> and acceleration fluctuations in a flat plate, zero pressure <span class="hlt">gradient</span> turbulent boundary layer at a momentum thickness Reynolds number of 7500 are measured using a spatially resolving three-component laser Doppler velocimeter. This canonical case data is applied for validation of the spectral approach for future application in more complex aerodynamic flows.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21443187-spider-iv-optical-near-infrared-color-gradients-early-type-galaxies-new-insight-correlations-galaxy-properties','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21443187-spider-iv-optical-near-infrared-color-gradients-early-type-galaxies-new-insight-correlations-galaxy-properties"><span>SPIDER. IV. OPTICAL AND NEAR-INFRARED COLOR <span class="hlt">GRADIENTS</span> IN EARLY-TYPE GALAXIES: NEW INSIGHT INTO CORRELATIONS WITH GALAXY PROPERTIES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>La Barbera, F.; De Carvalho, R. R.; De La Rosa, I. G.</p> <p>2010-11-15</p> <p>We present an analysis of stellar population <span class="hlt">gradients</span> in 4546 early-type galaxies (ETGs) with photometry in grizYHJK along with optical spectroscopy. ETGs were selected as bulge-dominated systems, displaying passive spectra within the SDSS fibers. A new approach is described which utilizes color information to constrain age and metallicity <span class="hlt">gradients</span>. Defining an effective color <span class="hlt">gradient</span>, {nabla}{sub *}, which incorporates all of the available color indices, we investigate how {nabla}{sub *} varies with galaxy mass proxies, i.e., <span class="hlt">velocity</span> dispersion, stellar (M{sub *}) and dynamical (M{sub dyn}) masses, as well as age, metallicity, and [{alpha}/Fe]. ETGs with M{sub dyn} larger than 8.5 xmore » 10{sup 10} M{sub sun} have increasing age <span class="hlt">gradients</span> and decreasing metallicity <span class="hlt">gradients</span> with respect to mass, metallicity, and enhancement. We find that <span class="hlt">velocity</span> dispersion and [{alpha}/Fe] are the main drivers of these correlations. ETGs with 2.5 x 10{sup 10} M{sub sun} {<=} M{sub dyn} {<=} 8.5 x 10{sup 10} M{sub sun} show no correlation of age, metallicity, and color <span class="hlt">gradients</span> with respect to mass, although color <span class="hlt">gradients</span> still correlate with stellar population parameters, and these correlations are independent of each other. In both mass regimes, the striking anti-correlation between color <span class="hlt">gradient</span> and {alpha}-enhancement is significant at {approx}5{sigma} and results from the fact that metallicity <span class="hlt">gradient</span> decreases with [{alpha}/Fe]. This anti-correlation may reflect the fact that star formation and metallicity enrichment are regulated by the interplay between the energy input from supernovae, and the temperature and pressure of the hot X-ray gas in ETGs. For all mass ranges, positive age <span class="hlt">gradients</span> are associated with old galaxies (>5-7 Gyr). For galaxies younger than {approx}5 Gyr, mostly at low mass, the age <span class="hlt">gradient</span> tends to be anti-correlated with the Age parameter, with more positive <span class="hlt">gradients</span> at younger ages.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...844L...2L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...844L...2L"><span>The Origins of UV-optical Color <span class="hlt">Gradients</span> in Star-forming Galaxies at z ˜ 2: Predominant Dust <span class="hlt">Gradients</span> but Negligible sSFR <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, F. S.; Jiang, Dongfei; Faber, S. M.; Koo, David C.; Yesuf, Hassen M.; Tacchella, Sandro; Mao, Shude; Wang, Weichen; Guo, Yicheng; Fang, Jerome J.; Barro, Guillermo; Zheng, Xianzhong; Jia, Meng; Tong, Wei; Liu, Lu; Meng, Xianmin</p> <p>2017-07-01</p> <p>The rest-frame UV-optical (I.e., NUV - B) color is sensitive to both low-level recent star formation (specific star formation rate—sSFR) and dust. In this Letter, we extend our previous work on the origins of NUV - B color <span class="hlt">gradients</span> in star-forming galaxies (SFGs) at z˜ 1 to those at z˜ 2. We use a sample of 1335 large (semimajor axis radius {R}{SMA}> 0\\buildrel{\\prime\\prime}\\over{.} 18) SFGs with extended UV emission out to 2{R}{SMA} in the mass range {M}* ={10}9{--}{10}11 {M}⊙ at 1.5< z< 2.8 in the CANDELS/GOODS-S and UDS fields. We show that these SFGs generally have negative NUV - B color <span class="hlt">gradients</span> (redder centers), and their color <span class="hlt">gradients</span> strongly increase with galaxy mass. We also show that the global rest-frame FUV - NUV color is approximately linear with {A}{{V}}, which is derived by modeling the <span class="hlt">observed</span> integrated FUV to NIR spectral energy distributions of the galaxies. Applying this integrated calibration to our spatially resolved data, we find a negative dust <span class="hlt">gradient</span> (more dust extinguished in the centers), which steadily becomes steeper with galaxy mass. We further find that the NUV - B color <span class="hlt">gradients</span> become nearly zero after correcting for dust <span class="hlt">gradients</span> regardless of galaxy mass. This indicates that the sSFR <span class="hlt">gradients</span> are negligible and dust reddening is likely the principal cause of negative UV-optical color <span class="hlt">gradients</span> in these SFGs. Our findings support that the buildup of the stellar mass in SFGs at Cosmic Noon is self-similar inside 2{R}{SMA}.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.G53A1124G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.G53A1124G"><span>Present-Day Strain and Rotation in the Lebanese Restraining Bend of the Dead Sea Fault System Based on Analysis of GPS <span class="hlt">Velocities</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gomez, F.; Jaafar, R.; Abdallah, C.; Karam, G.</p> <p>2012-12-01</p> <p>The Lebanese Restraining Bend (LRB) is a ~200-km-long bend in the central part of the Dead Sea Fault system (DSFS). As with other large restraining bends, this part of the transform is characterized by more complicated structure than other parts. Additionally, results from recent GPS studies have documented slower <span class="hlt">velocities</span> north of the LRB than are <span class="hlt">observed</span> along the southern DSFS to the south. In an effort to understand how strain is transferred through the LRB, this study analyzes improved GPS <span class="hlt">velocities</span> within the central DSFS based on new data and additional stations. Despite relatively modest rates of seismicity, the Dead Sea Fault system (DSFS) has a historically documented record of producing large and devastating earthquakes. Hence, geodetic measurements of crustal deformation may provide key constraints on processes of strain accumulation that may not be evident in instrumentally recorded seismicity. Within the LRB, the transform splays into two prominent strike-slip faults: The through-going Yammouneh fault and the Serghaya fault. The latter appears to terminate in the Anti-Lebanon Mountains. Additionally, some oblique plate motion is accommodated by thrusting along the coast of Lebanon. This study used GPS <span class="hlt">observations</span> from survey-mode GPS sites, as well as continuous GPS stations in the region. In total, 22 GPS survey sites have been measured in Lebanon between 2002 and 2010, along with GPS data from the adjacent area. Elastic models are used for initial assessment of fault slip rates. Incorporating two major strike-slip faults, as well as an offshore thrust fault, this modeling suggests left-lateral slip rates of 3.8 mm/yr and 1.1 mm/yr for the Yammouneh and Serghaya faults, respectively. The GPS survey network has sufficient density for analyzing <span class="hlt">velocity</span> <span class="hlt">gradients</span> in an effort to quantify tectonic strains and rotations. The <span class="hlt">velocity</span> <span class="hlt">gradients</span> suggest that differential rotations play a role in accommodating some plate motion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29056784','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29056784"><span>Theory and <span class="hlt">observations</span> of upward field-aligned currents at the magnetopause boundary layer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wing, Simon; Johnson, Jay R</p> <p>2015-11-16</p> <p>The dependence of the upward field-aligned current density ( J ‖ ) at the dayside magnetopause boundary layer is well described by a simple analytic model based on a <span class="hlt">velocity</span> shear generator. A previous <span class="hlt">observational</span> survey confirmed that the scaling properties predicted by the analytical model are applicable between 11 and 17 MLT. We utilize the analytic model to predict field-aligned currents using solar wind and ionospheric parameters and compare with direct <span class="hlt">observations</span>. The calculated and <span class="hlt">observed</span> parallel currents are in excellent agreement, suggesting that the model may be useful to infer boundary layer structures. However, near noon, where <span class="hlt">velocity</span> shear is small, the kinetic pressure <span class="hlt">gradients</span> and thermal currents, which are not included in the model, could make a small but significant contribution to J ‖ . Excluding data from noon, our least squares fit returns log( J ‖,max_cal ) = (0.96 ± 0.04) log( J ‖_obs ) + (0.03 ± 0.01) where J ‖,max_cal = calculated J ‖,max and J ‖_obs = <span class="hlt">observed</span> J ‖ .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.S41A2732L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.S41A2732L"><span>Efficient realization of 3D joint inversion of seismic and magnetotelluric data with cross <span class="hlt">gradient</span> structure constraint</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luo, H.; Zhang, H.; Gao, J.</p> <p>2016-12-01</p> <p>Seismic and magnetotelluric (MT) imaging methods are generally used to characterize subsurface structures at various scales. The two methods are complementary to each other and the integration of them is helpful for more reliably determining the resistivity and <span class="hlt">velocity</span> models of the target region. Because of the difficulty in finding empirical relationship between resistivity and <span class="hlt">velocity</span> parameters, Gallardo and Meju [2003] proposed a joint inversion method enforcing resistivity and <span class="hlt">velocity</span> models consistent in structure, which is realized by minimizing cross <span class="hlt">gradients</span> between two models. However, it is extremely challenging to combine two different inversion systems together along with the cross <span class="hlt">gradient</span> constraints. For this reason, Gallardo [2007] proposed a joint inversion scheme that decouples the seismic and MT inversion systems by iteratively performing seismic and MT inversions as well as cross <span class="hlt">gradient</span> minimization separately. This scheme avoids the complexity of combining two different systems together but it suffers the issue of balancing between data fitting and structure constraint. In this study, we have developed a new joint inversion scheme that avoids the problem encountered by the scheme of Gallardo [2007]. In the new scheme, seismic and MT inversions are still separately performed but the cross <span class="hlt">gradient</span> minimization is also constrained by model perturbations from separate inversions. In this way, the new scheme still avoids the complexity of combining two different systems together and at the same time the balance between data fitting and structure consistency constraint can be enforced. We have tested our joint inversion algorithm for both 2D and 3D cases. Synthetic tests show that joint inversion better reconstructed the <span class="hlt">velocity</span> and resistivity models than separate inversions. Compared to separate inversions, joint inversion can remove artifacts in the resistivity model and can improve the resolution for deeper resistivity structures. We</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006SbMat.197.1723B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006SbMat.197.1723B"><span>Discontinuous <span class="hlt">gradient</span> differential equations and trajectories in the calculus of variations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bogaevskii, I. A.</p> <p>2006-12-01</p> <p>The concept of <span class="hlt">gradient</span> of smooth functions is generalized for their sums with concave functions. An existence, uniqueness, and continuous dependence theorem for increasing time is formulated and proved for solutions of an ordinary differential equation the right-hand side of which is the <span class="hlt">gradient</span> of the sum of a concave and a smooth function. With the use of this result a physically natural motion of particles, well defined even at discontinuities of the <span class="hlt">velocity</span> field, is constructed in the variational problem of the minimal mechanical action in a space of arbitrary dimension. For such a motion of particles in the plane all typical cases of the birth and the interaction of point clusters of positive mass are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080047101&hterms=vector+fields&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dvector%2Bfields','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080047101&hterms=vector+fields&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dvector%2Bfields"><span>The Local Stellar <span class="hlt">Velocity</span> Field via Vector Spherical Harmonics</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Markarov, V. V.; Murphy, D. W.</p> <p>2007-01-01</p> <p>We analyze the local field of stellar tangential <span class="hlt">velocities</span> for a sample of 42,339 nonbinary Hipparcos stars with accurate parallaxes, using a vector spherical harmonic formalism. We derive simple relations between the parameters of the classical linear model (Ogorodnikov-Milne) of the local systemic field and low-degree terms of the general vector harmonic decomposition. Taking advantage of these relationships, we determine the solar <span class="hlt">velocity</span> with respect to the local stars of (V(sub X), V(sub Y), V(sub Z)) (10.5, 18.5, 7.3) +/- 0.1 km s(exp -1) not corrected for the asymmetric drift with respect to the local standard of rest. If only stars more distant than 100 pc are considered, the peculiar solar motion is (V(sub X), V(sub Y), V(sub Z)) (9.9, 15.6, 6.9) +/- 0.2 km s(exp -1). The adverse effects of harmonic leakage, which occurs between the reflex solar motion represented by the three electric vector harmonics in the <span class="hlt">velocity</span> space and higher degree harmonics in the proper-motion space, are eliminated in our analysis by direct subtraction of the reflex solar <span class="hlt">velocity</span> in its tangential components for each star. The Oort parameters determined by a straightforward least-squares adjustment in vector spherical harmonics are A=14.0 +/- 1.4, B=13.1 +/- 1.2, K=1.1 +/- 1.8, and C=2.9 +/- 1.4 km s(exp -1) kpc(exp -1). The physical meaning and the implications of these parameters are discussed in the framework of a general linear model of the <span class="hlt">velocity</span> field. We find a few statistically significant higher degree harmonic terms that do not correspond to any parameters in the classical linear model. One of them, a third-degree electric harmonic, is tentatively explained as the response to a negative linear <span class="hlt">gradient</span> of rotation <span class="hlt">velocity</span> with distance from the Galactic plane, which we estimate at approximately -20 km s(exp -1) kpc(exp -1). A similar vertical <span class="hlt">gradient</span> of rotation <span class="hlt">velocity</span> has been detected for more distant stars representing the thick disk (z greater than 1 kpc</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AdSpR..61.1672P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AdSpR..61.1672P"><span>Comparison of vertical E × B drift <span class="hlt">velocities</span> and ground-based magnetometer <span class="hlt">observations</span> of DELTA H in the low latitude under geomagnetically disturbed conditions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prabhu, M.; Unnikrishnan, K.</p> <p>2018-04-01</p> <p>In the present work, we analyzed the daytime vertical E × B drift <span class="hlt">velocities</span> obtained from Jicamarca Unattended Long-term Ionosphere Atmosphere (JULIA) radar and ΔH component of geomagnetic field measured as the difference between the magnitudes of the horizontal (H) components between two magnetometers deployed at two different locations Jicamarca, and Piura in Peru for 22 geomagnetically disturbed events in which either SC has occurred or Dstmax < -50 nT during the period 2006-2011. The ΔH component of geomagnetic field is measured as the differences in the magnitudes of horizontal H component between magnetometer placed directly on the magnetic equator and one displaced 6-9° away. It will provide a direct measure of the daytime electrojet current, due to the eastward electric field. This will in turn gives the magnitude of vertical E × B drift <span class="hlt">velocity</span> in the F region. A positive correlation exists between peak values of daytime vertical E × B drift <span class="hlt">velocity</span> and peak value of ΔH for the three consecutive days of the events. It was <span class="hlt">observed</span> that 45% of the events have daytime vertical E × B drift <span class="hlt">velocity</span> peak in the magnitude range 10-20 m/s and 20-30 m/s and 20% have peak ΔH in the magnitude range 50-60 nT and 80-90 nT. It was <span class="hlt">observed</span> that the time of occurrence of the peak value of both the vertical E × B drift <span class="hlt">velocity</span> and the ΔH have a maximum (40%) probability in the same time range 11:00-13:00 LT. We also investigated the correlation between E × B drift <span class="hlt">velocity</span> and Dst index and the correlation between delta H and Dst index. A strong positive correlation is found between E × B drift and Dst index as well as between delta H and Dst Index. Three different techniques of data analysis - linear, polynomial (order 2), and polynomial (order 3) regression analysis were considered. The regression parameters in all the three cases were calculated using the Least Square Method (LSM), using the daytime vertical E × B drift <span class="hlt">velocity</span> and ΔH. A formula</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JTePh..57.1307K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JTePh..57.1307K"><span>Thermoconvective flow <span class="hlt">velocity</span> in a high-speed magnetofluid seal after it has stopped</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krakov, M. S.; Nikiforov, I. V.</p> <p>2012-09-01</p> <p>Convective flow is investigated in the high-speed (linear <span class="hlt">velocity</span> of the shaft seal is more than 1 m/s) magnetofluid shaft seal after it has been stopped. Magnetic fluid is preliminarily heated due to viscous friction in the moving seal. After the shaft has been stopped, nonuniform heated fluid remains under the action of a high-<span class="hlt">gradient</span> magnetic field. Numerical analysis has revealed that in this situation, intense thermomagnetic convection is initiated. The <span class="hlt">velocity</span> of magnetic fluid depends on its viscosity. For the fluid with viscosity of 2 × 10-4 m2/s the maximum flow <span class="hlt">velocity</span> within the volume of magnetic fluid with a characteristic size of 1 mm can attain a value of 10 m/s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.969a2116S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.969a2116S"><span><span class="hlt">Observation</span> of spin superfluidity: YIG magnetic films and beyond</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sonin, Edouard</p> <p>2018-03-01</p> <p>From topology of the order parameter of the magnon condensate <span class="hlt">observed</span> in yttrium-iron-garnet (YIG) magnetic films one must not expect energetic barriers making spin supercurrents metastable. But we show that some barriers of dynamical origin are possible nevertheless until the <span class="hlt">gradient</span> of the phase (angle of spin precession) does not exceed the critical value (analog of the Landau critical <span class="hlt">velocity</span> in superfluids). On the other hand, recently published claims of experimental detection of spin superfluidity in YIG films and antiferromagnets are not justified, and spin superfluidity in magnetically ordered solids has not yet been experimentally confirmed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24838252','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24838252"><span>Volumetric <span class="hlt">velocity</span> measurements in restricted geometries using spiral sampling: a phantom study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nilsson, Anders; Revstedt, Johan; Heiberg, Einar; Ståhlberg, Freddy; Bloch, Karin Markenroth</p> <p>2015-04-01</p> <p>The aim of this study was to evaluate the accuracy of maximum <span class="hlt">velocity</span> measurements using volumetric phase-contrast imaging with spiral readouts in a stenotic flow phantom. In a phantom model, maximum <span class="hlt">velocity</span>, flow, pressure <span class="hlt">gradient</span>, and streamline visualizations were evaluated using volumetric phase-contrast magnetic resonance imaging (MRI) with <span class="hlt">velocity</span> encoding in one (extending on current clinical practice) and three directions (for characterization of the flow field) using spiral readouts. Results of maximum <span class="hlt">velocity</span> and pressure drop were compared to computational fluid dynamics (CFD) simulations, as well as corresponding low-echo-time (TE) Cartesian data. Flow was compared to 2D through-plane phase contrast (PC) upstream from the restriction. Results obtained with 3D through-plane PC as well as 4D PC at shortest TE using a spiral readout showed excellent agreements with the maximum <span class="hlt">velocity</span> values obtained with CFD (<1 % for both methods), while larger deviations were seen using Cartesian readouts (-2.3 and 13 %, respectively). Peak pressure drop calculations from 3D through-plane PC and 4D PC spiral sequences were respectively 14 and 13 % overestimated compared to CFD. Identification of the maximum <span class="hlt">velocity</span> location, as well as the accurate <span class="hlt">velocity</span> quantification can be obtained in stenotic regions using short-TE spiral volumetric PC imaging.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990100883','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990100883"><span>Influence of Applied Thermal <span class="hlt">Gradients</span> and a Static Magnetic Field on Bridgman-Grown GeSi Alloys</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Volz, M. P.; Szofran, F. R.; Cobb, S. D.; Ritter, T. M.</p> <p>1999-01-01</p> <p>The effect of applied axial and radial thermal <span class="hlt">gradients</span> and an axial static magnetic field on the macrosegregation profiles of Bridgman-grown GeSi alloy crystals has been assessed. The axial thermal <span class="hlt">gradients</span> were adjusted by changing the control setpoints of a seven-zone vertical Bridgman furnace. The radial thermal <span class="hlt">gradients</span> were affected by growing samples in ampoules with different thermal conductivities, namely graphite, hot-pressed boron nitride (BN), and pyrolytic boron nitride (PBN). Those samples grown in a graphite ampoule exhibited radial profiles consistent with a highly concave interface and axial profiles indicative of complete mixing in the melt. The samples grown in BN and PBN ampoules had less radial variation. Axial macrosegregation profiles of these samples fell between the predictions for a completely mixed melt and one where solute transport is dominated by diffusion. All of the samples were grown on Ge seeds. This resulted in a period of free growth until the Si concentration in the solid was in equilibrium with the Si concentration in the liquid. The length of crystal grown during this period was inversely proportional to the applied axial thermal <span class="hlt">gradient</span>. Several samples were grown in an axial 5 Tesla magnetic field. Measured macroscopic segregation profiles on these samples indicate that the magnetic field did not, in general, reduce the melt flow <span class="hlt">velocities</span> to below the growth <span class="hlt">velocities</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9647721','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9647721"><span><span class="hlt">Observation</span> of TiF+ by <span class="hlt">Velocity</span> Modulation Laser Spectroscopy and Analysis of the</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Focsa; Pinchemel; Collet; Huet</p> <p>1998-06-01</p> <p>The molecular ion TiF+ has been <span class="hlt">observed</span> for the first time using high-resolution spectroscopy. The ions were produced in the positive column of an AC glow discharge with a gas mixture of He/TiF4. A single-mode cw dye laser along with the <span class="hlt">velocity</span> modulation detection technique was used to record an absorption spectrum in the spectral region 16 800-18 600 cm-1. The <span class="hlt">observed</span> system was assigned to the 0-0 and 1-1 bands of the [17.6]3Delta-X3Phi transition of TiF+. The rotational analysis of the main subbands has been performed up to J values equal to 77 and 56 for the 0-0 and 1-1 bands, respectively. Despite a careful search, no intercombination band was <span class="hlt">observed</span>. A set of effective molecular parameters has been determined, characterizing the v = 0, 1 levels of the [17.6]3Delta and X3Phi states. The spin-orbit constants Ae and the vibrational constants omegae, omegaexe have been estimated for both electronic states, as well as their equilibrium distances Re (1.7509 and 1.7800) Å for the [17.6]3Delta and X3Phi states, respectively). Copyright 1998 Academic Press.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AAS...21933601S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AAS...21933601S"><span>The 6dFGS Peculiar <span class="hlt">Velocity</span> Field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Springob, Chris M.; Magoulas, C.; Colless, M.; Mould, J.; Erdogdu, P.; Jones, D. H.; Lucey, J.; Campbell, L.; Merson, A.; Jarrett, T.</p> <p>2012-01-01</p> <p>The 6dF Galaxy Survey (6dFGS) is an all southern sky galaxy survey, including 125,000 redshifts and a Fundamental Plane (FP) subsample of 10,000 peculiar <span class="hlt">velocities</span>, making it the largest peculiar <span class="hlt">velocity</span> sample to date. We have fit the FP using a maximum likelihood fit to a tri-variate Gaussian. We subsequently compute a Bayesian probability distribution for every possible peculiar <span class="hlt">velocity</span> for each of the 10,000 galaxies, derived from the tri-variate Gaussian probability density distribution, accounting for our selection effects and measurement errors. We construct a predicted peculiar <span class="hlt">velocity</span> field from the 2MASS redshift survey, and compare our <span class="hlt">observed</span> 6dFGS <span class="hlt">velocity</span> field to the predicted field. We discuss the resulting agreement between the <span class="hlt">observed</span> and predicted fields, and the implications for measurements of the bias parameter and bulk flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARE16005P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARE16005P"><span>Temperature-<span class="hlt">gradient</span>-induced</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Park, Cheol; Glaser, Matt; Maclennan, Joe; Clark, Noel; Trittel, Torsten; Stannarius, Ralf</p> <p></p> <p>Freely-suspended smectic films of sub-micrometer thickness and lateral extensions of several millimeters were used to study thermally driven migration and convection in the film plane. Film experiments were performed during the 6 minute microgravity phase of a TEXUS suborbital rocket flight (Texus 52, launched April 27, 2015). We have found an attraction of the smectic material towards the cold edge of the film in a temperature <span class="hlt">gradient</span>, similar to the Soret effect. This process is reversed when this edge is heated up again. Thermal convection driven by two thermocontacts in the film is practically absent, even at temperature <span class="hlt">gradients</span> up to 10 K/mm, with thermally driven convection only setting in when the hot post reaches the transition temperature to the nematic phase. The <span class="hlt">Observation</span> and Analysis of Smectic Islands in Space (OASIS) flight hardware was launched on SpaceX-6 in April 2015 and experiments on smectic bubbles were carried out on the International Space Station using four different smectic A and C liquid crystal materials in separate sample chambers. We <span class="hlt">observed</span> that smectic islands on the surface of the bubbles migrated towards the colder part of the bubble in a temperature <span class="hlt">gradient</span>. This work was supported by NASA Grant No. NNX-13AQ81G, by the Soft Materials Research Center under NSF MRSEC Grants No. DMR-0820579 and No. DMR-1420736, and by DLR Grants 50WM1127 and 50WM1430.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22654315-ray-radio-study-varying-expansion-velocities-tychos-supernova-remnant','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22654315-ray-radio-study-varying-expansion-velocities-tychos-supernova-remnant"><span>AN X-RAY AND RADIO STUDY OF THE VARYING EXPANSION <span class="hlt">VELOCITIES</span> IN TYCHO’S SUPERNOVA REMNANT</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Williams, Brian J.; Chomiuk, Laura; Hewitt, John W.</p> <p>2016-06-01</p> <p>We present newly obtained X-ray and radio <span class="hlt">observations</span> of Tycho’s supernova remnant using Chandra and the Karl G. Jansky Very Large Array in 2015 and 2013/14, respectively. When combined with earlier epoch <span class="hlt">observations</span> by these instruments, we now have time baselines for expansion measurements of the remnant of 12–15 years in the X-rays and 30 years in the radio. The remnant’s large angular size allows for proper motion measurements at many locations around the periphery of the blast wave. Consistent with earlier measurements, we find a clear <span class="hlt">gradient</span> in the expansion <span class="hlt">velocity</span> of the remnant, despite its round shape. Themore » proper motions on the western and southwestern sides of the remnant are about a factor of two higher than those in the east and northeast. We showed in an earlier work that this is related to an offset of the explosion site from the geometric center of the remnant due to a density <span class="hlt">gradient</span> in the ISM, and using our refined measurements reported here, we find that this offset is ∼23″ toward the northeast. An explosion center offset in such a circular remnant has implications for searches for progenitor companions in other remnants.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007APS..MARB19003A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007APS..MARB19003A"><span>Concentration and <span class="hlt">Velocity</span> Measurements of Both Phases in Liquid-Solid Slurries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Altobelli, Stephen; Hill, Kimberly; Caprihan, Arvind</p> <p>2007-03-01</p> <p>Natural and industrial slurry flows abound. They are difficult to calculate and to measure. We demonstrate a simple technique for studying steady slurries. We previously used time-of-flight techniques to study pressure driven slurry flow in pipes. Only the continuous phase <span class="hlt">velocity</span> and concentration fields were measured. The discrete phase concentration was inferred. In slurries composed of spherical, oil-filled pills and poly-methyl-siloxane oils, we were able to use inversion nulling to measure the concentration and <span class="hlt">velocity</span> fields of both phases. Pills are available in 1-5mm diameter and silicone oils are available in a wide range of viscosities, so a range of flows can be studied. We demonstrated the technique in horizontal, rotating cylinder flows. We combined two tried and true methods to do these experiments. The first used the difference in T1 to select between phases. The second used <span class="hlt">gradient</span> waveforms with controlled first moments to produce <span class="hlt">velocity</span> dependent phase shifts. One novel processing method was developed that allows us to use static continuous phase measurements to reference both the continuous and discrete phase <span class="hlt">velocity</span> images. ?</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.9566J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.9566J"><span>Lithospheric Structure of Arabia from the Joint Inversion of P- and S-wave Receiver Functions and Dispersion <span class="hlt">Velocities</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Julia, Jordi; Al-Amri, Abdullah; Pasyanos, Michael; Rodgers, Arthur; Matzel, Eric; Nyblade, Andrew</p> <p>2013-04-01</p> <p>Seismic imaging of the lithosphere under the Arabian shield and platform is critical to help answer important geologic questions of regional and global interest. The Arabian Shield can be regarded as an amalgamation of several arcs and microplates of Proterozoic age that culminated in the accretion of the Arabian portion of Gondwana during the Pan-African event at ~550 Ma and the role of important geologic features <span class="hlt">observed</span> on the surface - such as the lineaments and shear zones separating the Proterozoic terrains in the shield - is not completely understood. Also, current models of Precambrian crustal evolution predict that Proterozoic terranes are underlain by fertile (FeO-rich) cratonic roots that should promote the production of mafic magmas and underplating of the Arabian shield terranes, and the shield contains Tertiary and Quaternary volcanic rocks related to the early stages of the Red Sea formation that might also be related to plume-related lithospheric "erosion". In order to better understand these relationships, we are developing new <span class="hlt">velocity</span> models of litospheric structure for the Arabian shield and platform from the joint inversion of up to four seismic data sets: P-wave receiver functions, S-wave receiver functions, dispersion <span class="hlt">velocities</span> from surface-waves, and dispersion <span class="hlt">velocities</span> from ambient-noise cross-correlations. The joint inversion combines constraints on crustal thickness from P-wave receiver functions, constraints on lithospheric thickness from S-wave receiver functions and constraints on S-<span class="hlt">velocity</span> and S-<span class="hlt">velocity</span> <span class="hlt">gradients</span> from dispersion <span class="hlt">velocities</span> to produce detailed S-<span class="hlt">velocity</span> profiles under single recording stations. We will present S-<span class="hlt">velocity</span> profiles for a number of permanent stations operated by the Saudi Geological Survey and the King ing Abdulaziz Center for Science and Technology as well as stations from past temporary deployments and discuss the implications of the <span class="hlt">velocity</span> models regarding composition and tectonics of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997GeoRL..24....9R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997GeoRL..24....9R"><span>Low crustal <span class="hlt">velocities</span> and mantle lithospheric variations in southern Tibet from regional Pnl waveforms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodgers, Arthur J.; Schwartz, Susan Y.</p> <p></p> <p>We report low average crustal P-wave <span class="hlt">velocities</span> (5.9-6.1 km/s, Poisson's ratio 0.23-0.27, thickness 68-76 km) in southern Tibet from modelling regional Pnl waveforms recorded by the 1991-1992 Tibetan Plateau Experiment. We also find that the mantle lithosphere beneath the Indus-Tsangpo Suture and the Lhasa Terrane is shield-like (Pn <span class="hlt">velocity</span> 8.20-8.25 km/s, lid thickness 80-140 km, positive <span class="hlt">velocity</span> <span class="hlt">gradient</span> 0.0015-0.0025 s-1). Analysis of relative Pn travel time residuals requires a decrease in the mantle <span class="hlt">velocities</span> beneath the northern Lhasa Terrane, the Banggong-Nujiang Suture and the southern Qiangtang Terrane. Tectonic and petrologic considerations suggest that low bulk crustal <span class="hlt">velocities</span> could result from a thick (50-60 km) felsic upper crust with vertically limited and laterally pervasive partial melt. These results are consistent with underthrusting of Indian Shield lithosphere beneath the Tibetan Plateau to at least the central Lhasa Terrane.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMEP43C0975T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP43C0975T"><span>Influence of Microclimate on Semi-Arid Montane Conifer Forest Sapflux <span class="hlt">Velocity</span> in Complex Terrain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thirouin, K. R.; Barnard, D. M.; Barnard, H. R.</p> <p>2016-12-01</p> <p>Microclimate variation in complex terrain is key to our understanding of large-scale climate change effects on montane ecosystems. Modern climate models forecast that semi-arid montane ecosystems in the western United States are to experience increases in temperature, number of extreme drought events, and decreases in annual snowpack, all of which will potentially influence ecosystem water, carbon, and energy balances. In this study, we developed response curves that describe the relationships between stem sapflux <span class="hlt">velocity</span>, air temperature (Tair), incoming solar radiation (SWin), soil temperature (Tsoil), and soil moisture content (VWC) in sites of Pinus contorta and Pinus ponderosa distributed along an elevation and aspect <span class="hlt">gradient</span> in the montane zone of the Central Rocky Mountains, Colorado, USA. Among sites we found sapflux <span class="hlt">velocity</span> to be significantly correlated with all four environmental factors (p < 0.05), but most strongly with SWin and Tair. The response of sapflux <span class="hlt">velocity</span> to SWin was logarithmic, whereas the response to Tair indicated a peak sapflux <span class="hlt">velocity</span> at a site-specific temperature that declined with increasing Tair. Sapflux <span class="hlt">velocity</span> also increased with increasing VWC, but decreased with increasing Tsoil. At south-facing sites, the initial increase in the logarithmic response curve for SWin leveled off at 150-250 W m-2, whereas for north-facing sites it leveled off at 50-125 W m-2. While the differences in the SWin response between aspects could be due to species physiological differences, the highest elevation south-facing P. contorta site behaved similarly to the south-facing P. ponderosa, suggesting that environmental drivers may dominate the response. In response to Tair, peak sapflux <span class="hlt">velocity</span> occurred at 12-13 degrees C at all sites except the mid-slope north-facing P. contorta site, which also had the lowest Tsoil. The responses of stem sapflux <span class="hlt">velocity</span> to climate drivers indicate that forest transpiration is regulated by microclimate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22286209','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22286209"><span>Nanofiber scaffold <span class="hlt">gradients</span> for interfacial tissue engineering.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ramalingam, Murugan; Young, Marian F; Thomas, Vinoy; Sun, Limin; Chow, Laurence C; Tison, Christopher K; Chatterjee, Kaushik; Miles, William C; Simon, Carl G</p> <p>2013-02-01</p> <p>We have designed a 2-spinnerette device that can directly electrospin nanofiber scaffolds containing a <span class="hlt">gradient</span> in composition that can be used to engineer interfacial tissues such as ligament and tendon. Two types of nanofibers are simultaneously electrospun in an overlapping pattern to create a nonwoven mat of nanofibers containing a composition <span class="hlt">gradient</span>. The approach is an advance over previous methods due to its versatility - <span class="hlt">gradients</span> can be formed from any materials that can be electrospun. A dye was used to characterize the 2-spinnerette approach and applicability to tissue engineering was demonstrated by fabricating nanofibers with <span class="hlt">gradients</span> in amorphous calcium phosphate nanoparticles (nACP). Adhesion and proliferation of osteogenic cells (MC3T3-E1 murine pre-osteoblasts) on <span class="hlt">gradients</span> was enhanced on the regions of the <span class="hlt">gradients</span> that contained higher nACP content yielding a graded osteoblast response. Since increases in soluble calcium and phosphate ions stimulate osteoblast function, we measured their release and <span class="hlt">observed</span> significant release from nanofibers containing nACP. The nanofiber-nACP <span class="hlt">gradients</span> fabricated herein can be applied to generate tissues with osteoblast <span class="hlt">gradients</span> such as ligaments or tendons. In conclusion, these results introduce a versatile approach for fabricating nanofiber <span class="hlt">gradients</span> that can have application for engineering graded tissues.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820048888&hterms=cholesterol&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcholesterol','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820048888&hterms=cholesterol&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcholesterol"><span><span class="hlt">Velocity</span> and attenuation of sound in arterial tissues</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rooney, J. A.; Gammell, P. M.; Hestenes, J. D.; Chin, H. P.; Blankenhorn, D. H.</p> <p>1982-01-01</p> <p>The <span class="hlt">velocity</span> of sound in excised human and canine arterial tissues is measured in order to serve as a basis for the development and application of ultrasonic techniques for the diagnosis of atherosclerotic lesions. Measurements of sound <span class="hlt">velocity</span> at different regions of 11 human and six canine aortas were made by a time delay spectrometer technique at frequencies from 2 to 10 MHz, and compared with ultrasonic attenuation parameters and the results of biochemical assays. Sound <span class="hlt">velocity</span> is found to increase with increasing attenuation at all frequencies, and with increasing collagen content. A strong dependence of sound <span class="hlt">velocity</span> on cholesterol content or low calcium contents is not <span class="hlt">observed</span>, although <span class="hlt">velocities</span> of up to 2000 m/sec are <span class="hlt">observed</span> in highly organized calcified lesions. A decrease in <span class="hlt">velocity</span> with decreasing temperature is also noted. It is thus concluded that it is principally the differences in tissue collagen levels that contribute to image formation according to sound <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhTea..53..360D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhTea..53..360D"><span>Demonstrating the Direction of Angular <span class="hlt">Velocity</span> in Circular Motion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Demircioglu, Salih; Yurumezoglu, Kemal; Isik, Hakan</p> <p>2015-09-01</p> <p>Rotational motion is ubiquitous in nature, from astronomical systems to household devices in everyday life to elementary models of atoms. Unlike the tangential <span class="hlt">velocity</span> vector that represents the instantaneous linear <span class="hlt">velocity</span> (magnitude and direction), an angular <span class="hlt">velocity</span> vector is conceptually more challenging for students to grasp. In physics classrooms, the direction of an angular <span class="hlt">velocity</span> vector is taught by the right-hand rule, a mnemonic tool intended to aid memory. A setup constructed for instructional purposes may provide students with a more easily understood and concrete method to <span class="hlt">observe</span> the direction of the angular <span class="hlt">velocity</span>. This article attempts to demonstrate the angular <span class="hlt">velocity</span> vector using the <span class="hlt">observable</span> motion of a screw mounted to a remotely operated toy car.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005JFM...524..143A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005JFM...524..143A"><span>Studies of interactions of a propagating shock wave with decaying grid turbulence: <span class="hlt">velocity</span> and vorticity fields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Agui, Juan H.; Briassulis, George; Andreopoulos, Yiannis</p> <p>2005-02-01</p> <p>The unsteady interaction of a moving shock wave with nearly homogeneous and isotropic decaying compressible turbulence has been studied experimentally in a large-scale shock tube facility. Rectangular grids of various mesh sizes were used to generate turbulence with Reynolds numbers based on Taylor's microscale ranging from 260 to 1300. The interaction has been investigated by measuring the three-dimensional <span class="hlt">velocity</span> and vorticity vectors, the full <span class="hlt">velocity</span> <span class="hlt">gradient</span> and rate-of-strain tensors with instrumentation of high temporal and spatial resolution. This allowed estimates of dilatation, compressible dissipation and dilatational stretching to be obtained. The time-dependent signals of enstrophy, vortex stretching/tilting vector and dilatational stretching vector were found to exhibit a rather strong intermittent behaviour which is characterized by high-amplitude bursts with values up to 8 times their r.m.s. within periods of less violent and longer lived events. Several of these bursts are evident in all the signals, suggesting the existence of a dynamical flow phenomenon as a common cause. Fluctuations of all <span class="hlt">velocity</span> <span class="hlt">gradients</span> in the longitudinal direction are amplified significantly downstream of the interaction. Fluctuations of the <span class="hlt">velocity</span> <span class="hlt">gradients</span> in the lateral directions show no change or a minor reduction through the interaction. Root mean square values of the lateral vorticity components indicate a 25% amplification on average, which appears to be very weakly dependent on the shock strength. The transmission of the longitudinal vorticity fluctuations through the shock appears to be less affected by the interaction than the fluctuations of the lateral components. Non-dissipative vortex tubes and irrotational dissipative motions are more intense in the region downstream of the shock. There is also a significant increase in the number of events with intense rotational and dissipative motions. Integral length scales and Taylor's microscales were reduced</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApGeo..14..381S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApGeo..14..381S"><span>Conjugate <span class="hlt">gradient</span> and cross-correlation based least-square reverse time migration and its application</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Xiao-Dong; Ge, Zhong-Hui; Li, Zhen-Chun</p> <p>2017-09-01</p> <p>Although conventional reverse time migration can be perfectly applied to structural imaging it lacks the capability of enabling detailed delineation of a lithological reservoir due to irregular illumination. To obtain reliable reflectivity of the subsurface it is necessary to solve the imaging problem using inversion. The least-square reverse time migration (LSRTM) (also known as linearized reflectivity inversion) aims to obtain relatively high-resolution amplitude preserving imaging by including the inverse of the Hessian matrix. In practice, the conjugate <span class="hlt">gradient</span> algorithm is proven to be an efficient iterative method for enabling use of LSRTM. The <span class="hlt">velocity</span> <span class="hlt">gradient</span> can be derived from a cross-correlation between <span class="hlt">observed</span> data and simulated data, making LSRTM independent of wavelet signature and thus more robust in practice. Tests on synthetic and marine data show that LSRTM has good potential for use in reservoir description and four-dimensional (4D) seismic images compared to traditional RTM and Fourier finite difference (FFD) migration. This paper investigates the first order approximation of LSRTM, which is also known as the linear Born approximation. However, for more complex geological structures a higher order approximation should be considered to improve imaging quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017BoLMe.165..385K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017BoLMe.165..385K"><span>A Modulated-<span class="hlt">Gradient</span> Parametrization for the Large-Eddy Simulation of the Atmospheric Boundary Layer Using the Weather Research and Forecasting Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khani, Sina; Porté-Agel, Fernando</p> <p>2017-12-01</p> <p>The performance of the modulated-<span class="hlt">gradient</span> subgrid-scale (SGS) model is investigated using large-eddy simulation (LES) of the neutral atmospheric boundary layer within the weather research and forecasting model. Since the model includes a finite-difference scheme for spatial derivatives, the discretization errors may affect the simulation results. We focus here on understanding the effects of finite-difference schemes on the momentum balance and the mean <span class="hlt">velocity</span> distribution, and the requirement (or not) of the ad hoc canopy model. We find that, unlike the Smagorinsky and turbulent kinetic energy (TKE) models, the calculated mean <span class="hlt">velocity</span> and vertical shear using the modulated-<span class="hlt">gradient</span> model, are in good agreement with Monin-Obukhov similarity theory, without the need for an extra near-wall canopy model. The structure of the near-wall turbulent eddies is better resolved using the modulated-<span class="hlt">gradient</span> model in comparison with the classical Smagorinsky and TKE models, which are too dissipative and yield unrealistic smoothing of the smallest resolved scales. Moreover, the SGS fluxes obtained from the modulated-<span class="hlt">gradient</span> model are much smaller near the wall in comparison with those obtained from the regular Smagorinsky and TKE models. The apparent inability of the LES model in reproducing the mean streamwise component of the momentum balance using the total (resolved plus SGS) stress near the surface is probably due to the effect of the discretization errors, which can be calculated a posteriori using the Taylor-series expansion of the resolved <span class="hlt">velocity</span> field. Overall, we demonstrate that the modulated-<span class="hlt">gradient</span> model is less dissipative and yields more accurate results in comparison with the classical Smagorinsky model, with similar computational costs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DPPN11141S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DPPN11141S"><span>Hydrodynamic Model for Density <span class="hlt">Gradients</span> Instability in Hall Plasmas Thrusters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Sukhmander</p> <p>2017-10-01</p> <p>There is an increasing interest for a correct understanding of purely growing electromagnetic and electrostatic instabilities driven by a plasma <span class="hlt">gradient</span> in a Hall thruster devices. In Hall thrusters, which are typically operated with xenon, the thrust is provided by the acceleration of ions in the plasma generated in a discharge chamber. The goal of this paper is to study the instabilities due to <span class="hlt">gradients</span> of plasma density and conditions for the growth rate and real part of the frequency for Hall thruster plasmas. Inhomogeneous plasmas prone a wide class of eigen modes induced by inhomogeneities of plasma density and called drift waves and instabilities. The growth rate of the instability has a dependences on the magnetic field, plasma density, ion temperature and wave numbers and initial drift <span class="hlt">velocities</span> of the plasma species.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22472189-influence-radiation-absorption-microparticles-flame-velocity-combustion-regimes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22472189-influence-radiation-absorption-microparticles-flame-velocity-combustion-regimes"><span>Influence of radiation absorption by microparticles on the flame <span class="hlt">velocity</span> and combustion regimes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ivanov, M. F., E-mail: ivanov-mf@mail.ru; Kiverin, A. D.; Liberman, M. A., E-mail: michael.liberman@nordita.org</p> <p></p> <p>Thermal radiation from hot combustion products has virtually no effect on the flame propagation in a gas medium. We consider a different situation when even a small concentration of microparticles suspended in a gas absorbs the thermal radiation and heats the gas mixture ahead of the combustion wave front by transferring it to the gas. The mixture heating ahead of the flame front can lead either to a moderate increase in the combustion wave <span class="hlt">velocity</span> for a fast flame or to its significant increase for a slow flame, depending on the gas mixture reactivity and the normal laminar flame <span class="hlt">velocity</span>.more » For a slow flame, the heat transfer by radiation from the combustion products can become the dominant mechanism compared to the ordinary molecular thermal conduction that determines the combustion wave structure and <span class="hlt">velocity</span>. The radiative heating for a spatially nonuniform distribution of particles ahead of the flame front is shown to give rise to a temperature <span class="hlt">gradient</span> that, in turn, can lead to the ignition of different combustion regimes, depending on the radiation absorption length. In accordance with the Zeldovich <span class="hlt">gradient</span> mechanism, both deflagration and detonation regimes can be formed in this case. A hydrogen–oxygen flame is used as an example to illustrate the ignition of different combustion wave propagation regimes, depending on the radiation absorption length.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3880383','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3880383"><span>A pressure-<span class="hlt">gradient</span> mechanism for vortex shedding in constricted channels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Boghosian, M. E.; Cassel, K. W.</p> <p>2013-01-01</p> <p>Numerical simulations of the unsteady, two-dimensional, incompressible Navier–Stokes equations are performed for a Newtonian fluid in a channel having a symmetric constriction modeled by a two-parameter Gaussian distribution on both channel walls. The Reynolds number based on inlet half-channel height and mean inlet <span class="hlt">velocity</span> ranges from 1 to 3000. Constriction ratios based on the half-channel height of 0.25, 0.5, and 0.75 are considered. The results show that both the Reynolds number and constriction geometry have a significant effect on the behavior of the post-constriction flow field. The Navier–Stokes solutions are <span class="hlt">observed</span> to experience a number of bifurcations: steady attached flow, steady separated flow (symmetric and asymmetric), and unsteady vortex shedding downstream of the constriction depending on the Reynolds number and constriction ratio. A sequence of events is described showing how a sustained spatially growing flow instability, reminiscent of a convective instability, leads to the vortex shedding phenomenon via a proposed streamwise pressure-<span class="hlt">gradient</span> mechanism. PMID:24399860</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDL24001F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDL24001F"><span>Three-Dimensional <span class="hlt">Velocity</span> Field De-Noising using Modal Projection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frank, Sarah; Ameli, Siavash; Szeri, Andrew; Shadden, Shawn</p> <p>2017-11-01</p> <p>PCMRI and Doppler ultrasound are common modalities for imaging <span class="hlt">velocity</span> fields inside the body (e.g. blood, air, etc) and PCMRI is increasingly being used for other fluid mechanics applications where optical imaging is difficult. This type of imaging is typically applied to internal flows, which are strongly influenced by domain geometry. While these technologies are evolving, it remains that measured data is noisy and boundary layers are poorly resolved. We have developed a boundary modal analysis method to de-noise 3D <span class="hlt">velocity</span> fields such that the resulting field is divergence-free and satisfies no-slip/no-penetration boundary conditions. First, two sets of divergence-free modes are computed based on domain geometry. The first set accounts for flow through ``truncation boundaries'', and the second set of modes has no-slip/no-penetration conditions imposed on all boundaries. The modes are calculated by minimizing the <span class="hlt">velocity</span> <span class="hlt">gradient</span> throughout the domain while enforcing a divergence-free condition. The measured <span class="hlt">velocity</span> field is then projected onto these modes using a least squares algorithm. This method is demonstrated on CFD simulations with artificial noise. Different degrees of noise and different numbers of modes are tested to reveal the capabilities of the approach. American Heart Association Award 17PRE33660202.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001643.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001643.html"><span><span class="hlt">Gradient</span> Sun [still</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>To view a video of the <span class="hlt">Gradient</span> Sun go to: www.flickr.com/photos/gsfc/8103212817 Looking at a particularly beautiful image of the sun helps show how the lines between science and art can sometimes blur. But there is more to the connection between the two disciplines: science and art techniques are often quite similar, indeed one may inform the other or be improved based on lessons from the other arena. One such case is a technique known as a "<span class="hlt">gradient</span> filter" – recognizable to many people as an option available on a photo-editing program. <span class="hlt">Gradients</span> are, in fact, a mathematical description that highlights the places of greatest physical change in space. A <span class="hlt">gradient</span> filter, in turn, enhances places of contrast, making them all the more obviously different, a useful tool when adjusting photos. Scientists, too, use <span class="hlt">gradient</span> filters to enhance contrast, using them to accentuate fine structures that might otherwise be lost in the background noise. On the sun, for example, scientists wish to study a phenomenon known as coronal loops, which are giant arcs of solar material constrained to travel along that particular path by the magnetic fields in the sun's atmosphere. <span class="hlt">Observations</span> of the loops, which can be more or less tangled and complex during different phases of the sun's 11-year activity cycle, can help researchers understand what's happening with the sun's complex magnetic fields, fields that can also power great eruptions on the sun such as solar flares or coronal mass ejections. The still here shows an unfiltered image from the sun next to one that has been processed using a <span class="hlt">gradient</span> filter. Note how the coronal loops are sharp and defined, making them all the more easy to study. On the other hand, <span class="hlt">gradients</span> also make great art. NASA/Goddard Space Flight Center To download this video go to: svs.gsfc.nasa.gov/goto?11112 NASA image use policy. NASA Goddard Space Flight Center enables NASA’s mission through four scientific endeavors: Earth Science, Heliophysics</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMDI51C0331L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMDI51C0331L"><span><span class="hlt">Velocity</span> structure around the 410 km discontinuity beneath the East China Sea based on the waveform modeling method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, W.; Cui, Q.; Gao, Y.; Wei, R.; Zhou, Y.; Yu, J.</p> <p>2017-12-01</p> <p>The 410 km discontinuity is the upper boundary of the mantle transition zone. Seismic detections on the structure and morphology of the 410 km discontinuity are helpful to understand the compositions of the Earth's interior and the relevant geodynamics. In this paper, we select the broadband P waveforms of an intermediate earthquake that occurred in the Ryukyu subduction zone and retrieved from the China Digital Seismograph Network, and study the fine <span class="hlt">velocity</span> structure around the 410 km discontinuity by matching the <span class="hlt">observed</span> triplicated waveforms with the theoretical ones. Our results reveal that (1) the 410 km discontinuity beneath the East China Sea is mostly a sharp boundary with a small-scale uplift of 8-15 km and a <span class="hlt">gradient</span> boundary up to 20 km in the most southern part, and (2) there exist a low <span class="hlt">velocity</span> layer atop the 410 km discontinuity with the thickness of 50-62 km and P-wave <span class="hlt">velocity</span> decrease of 0.5%-1.5%, and (3) a high <span class="hlt">velocity</span> anomaly with P-wave decrease of 1.0%-3.0% below 440 km. Combining with the previous topographic results in this area, we speculate that the high <span class="hlt">velocity</span> anomaly is relevant to the stagnancy of the western Pacific slab in the mantle transition zone, the decomposition of phase E in the slab results in the increase of water content, which would cause the uplift of the 410 km discontinuity, and the low <span class="hlt">velocity</span> layer atop the discontinuity should be related to the partial melting of the mantle peridotite induced by the dehydration of the hydrous minerals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014acm..conf..537T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014acm..conf..537T"><span>Experimental study on the ejecta-<span class="hlt">velocity</span> distributions caused by low-<span class="hlt">velocity</span> impacts on quartz sand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsujido, S.; Arakawa, M.; Suzuki, A. I.; Yasui, M.</p> <p>2014-07-01</p> <p>Introduction: Regolith formation on asteroids is caused by successive impacts of small bodies. The ejecta <span class="hlt">velocity</span> distribution during the crater formation process is one of the most important physical properties related to the surface-evolution process, and the distribution is also necessary to reconstruct the planetary-accretion process among planetesimals. The surface of small bodies, such as asteroids and planetesimals in the solar system, could have varying porosity, strength, and density, and the impact <span class="hlt">velocity</span> could vary across a wide range from a few tens of m/s to several km/s. Therefore, it is necessary to conduct impact experiments by changing the physical properties of the target and the projectile in a wide <span class="hlt">velocity</span> range in order to constrain the crater-formation process applicable to the small bodies in the solar system. Housen and Holsapple (2011) compiled the data of ejecta <span class="hlt">velocity</span> distribution with various impact <span class="hlt">velocities</span>, porosities, grain sizes, grain shapes, and strengths of the targets, and they improved their ejecta scaling law. But the ejecta <span class="hlt">velocity</span> data is not enough for varying projectile densities and for impact <span class="hlt">velocities</span> less than 1 km/s. In this study, to investigate the projectile density dependence of the ejecta <span class="hlt">velocity</span> distribution at a low <span class="hlt">velocity</span> region, we conducted impact experiments with projectile densities from 1.1 to 11.3 g/cm^3. Then, we try to determine the effect of projectile density on the ejecta <span class="hlt">velocity</span> distribution by means of the <span class="hlt">observation</span> of each individual ejecta grain. Experimental methods: We made impact cratering experiments by using a vertical-type one-stage light-gas gun (V-LGG) set at Kobe University. Targets were quartz sand (irregular shape) and glass beads (spherical shape) with the grain size of 500 μ m (porosity 44.7 %). The target container with the size of 30 cm was set in a large vacuum chamber with air pressure less than 10^3 Pa. The projectile materials that we used were lead, copper</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/836657','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/836657"><span>Improved alternating <span class="hlt">gradient</span> transport and focusing of neutral molecules</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kalnins, Juris; Lambertson, Glen; Gould, Harvey</p> <p>2001-12-02</p> <p>Polar molecules, in strong-field seeking states, can be transported and focused by an alternating sequence of electric field <span class="hlt">gradients</span> that focus in one transverse direction while defocusing in the other. We show by calculation and numerical simulation, how one may greatly improve the alternating <span class="hlt">gradient</span> transport and focusing of molecules. We use a new optimized multipole lens design, a FODO lattice beam transport line, and lenses to match the beam transport line to the beam source and the final focus. We derive analytic expressions for the potentials, fields, and <span class="hlt">gradients</span> that may be used to design these lenses. We describemore » a simple lens optimization procedure and derive the equations of motion for tracking molecules through a beam transport line. As an example, we model a straight beamline that transports a 560 m/s jet-source beam of methyl fluoride molecules 15 m from its source and focuses it to 2 mm diameter. We calculate the beam transport line acceptance and transmission, for a beam with <span class="hlt">velocity</span> spread, and estimate the transmitted intensity for specified source conditions. Possible applications are discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EPJP..133..170R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EPJP..133..170R"><span>Temperature and porosity effects on wave propagation in nanobeams using bi-Helmholtz nonlocal strain-<span class="hlt">gradient</span> elasticity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reza Barati, Mohammad</p> <p>2018-05-01</p> <p>In this paper, applying a general nonlocal strain-<span class="hlt">gradient</span> elasticity model with two nonlocal and one strain-<span class="hlt">gradient</span> parameters, wave dispersion behavior of thermally affected and elastically bonded nanobeams is investigated. The two nanobeams are considered to have material imperfections or porosities evenly dispersed across the thickness. Each nanobeam has uniform thickness and is modeled by refined shear deformation beam theory with sinusoidal transverse shear strains. The governing equations of the system are derived by Hamilton's rule and are analytically solved to obtain wave frequencies and the <span class="hlt">velocity</span> of wave propagation. In the presented graphs, one can see that porosities, temperature, nonlocal, strain <span class="hlt">gradient</span> and bonding springs have great influences on the wave characteristics of the system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMNH23A1860A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMNH23A1860A"><span>Detiding Tsunami Currents to Validate <span class="hlt">Velocities</span> in Numerical Simulation Codes using <span class="hlt">Observations</span> Near Hawaii from the 2011 Tohoku Tsunami</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Adams, L. M.; LeVeque, R. J.</p> <p>2015-12-01</p> <p>The ability to measure, predict, and compute tsunami flow <span class="hlt">velocities</span> is ofimportance in risk assessment and hazard mitigation. Until recently, fewdirect measurements of tsunami <span class="hlt">velocities</span> existed to compare with modelresults. During the 11 March 2001 Tohoku Tsunami, 328 current meters werewere in place around the Hawaiian Islands, USA, that captured time seriesof water <span class="hlt">velocity</span> in 18 locations, in both harbors and deep channels, ata series of depths. Arcos and LeVeque[1] compared these records againstnumerical simulations performed using the GeoClaw numerical tsunami modelwhich is based on the depth-averaged shallow water equations. They confirmedthat GeoClaw can accurately predict <span class="hlt">velocities</span> at nearshore locations, andthat tsunami current <span class="hlt">velocity</span> is more spatially variable than wave formor height and potentially more sensitive for model validation.We present a new approach to detiding this sensitive current data. Thisapproach can be used separately on data at each depth of a current gauge.When averaged across depths, the Geoclaw results in [1] are validated. Withoutaveraging, the results should be useful to researchers wishing to validate their3D codes. These results can be downloaded from the project website below.The approach decomposes the pre-tsunami component of the data into three parts:a tidal component, a fast component (noise), and a slow component (not matchedby the harmonic analysis). Each part is extended to the time when the tsunamiis present and subtracted from the current data then to give the ''tsunami current''that can be compared with 2D or 3D codes that do not model currents in thepre-tsunami regime. [1] "Validating <span class="hlt">Velocities</span> in the GeoClaw Tsunami Model using <span class="hlt">Observations</span> NearHawaii from the 2001 Tohoku Tsunami"M.E.M. Arcos and Randall J. LeVequearXiv:1410.2884v1 [physics.geo-py], 10 Oct. 2014.project website: http://faculty.washington.edu/lma3/research.html</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018577','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018577"><span>A revised <span class="hlt">velocity</span>-reversal and sediment-sorting model for a high-<span class="hlt">gradient</span>, pool-riffle stream</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thompson, D.M.; Wohl, E.E.; Jarrett, R.D.</p> <p>1996-01-01</p> <p>Sediment-sorting processes related to varying channel-bed morphology were investigated from April to November 1993 along a 1-km pool-riffle and step-pool reach of North Saint Vrain Creek, a small mountain stream in the Rocky Mountains of northern Colorado. Measured cross-sectional areas of flow were used to suggest higher <span class="hlt">velocities</span> in pools than in riffles at high flow. Three hundred and sixteen tracer particles, ranging in size from 16 mm to 256 mm, were placed in two separate pool-riffle-pool sequences and used to assess sediment-sorting patterns and sediment-transport competence variations. Tracer-particle depositional evidence indicated higher sediment-transport competence in pools than in riffles at high flow. Pool-riffle sediment sorting may be created by <span class="hlt">velocity</span> reversals, and more localized sorting results from gravitational forces along the upstream sloping portion of the channel bed located at the downstream end of pools.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003APS..DFD.KD009C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003APS..DFD.KD009C"><span>Convective Electrokinetic Instability With Conductivity <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Chuan-Hua; Lin, Hao; Lele, Sanjiva; Santiago, Juan</p> <p>2003-11-01</p> <p>Electrokinetic flow instability has been experimentally identified and quantified in a glass T-junction microchannel system with a cross section of 11 um x 155 um. In this system, buffers of different conductivities were electrokinetically driven into a common mixing channel by a DC electric field. A convective instability was <span class="hlt">observed</span> with a threshold electric field of 0.45 kV/cm for a 10:1 conductivity ratio. A physical model has been developed which consists of a modified Ohmic model formulation for electrolyte solutions and the Navier-Stokes equations with an electric body force term. The model and experiments show that bulk charge accumulation in regions of conductivity <span class="hlt">gradients</span> is the key mechanism of such instabilities. A linear stability analysis was performed in a convective framework, and Briggs-Bers criteria were applied to determine the nature of instability. The analysis shows the instability is governed by two key parameters: the ratio of molecular diffusion to electroviscous time scale which governs the onset of instability, and the ratio of electroviscous to electroosmotic <span class="hlt">velocity</span> which governs whether the instability is convective or absolute. The model predicted critical electric field, growth rate, wavelength, and phase speed which were comparable to experimental data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/541187-calibration-method-helps-seismic-velocity-interpretation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/541187-calibration-method-helps-seismic-velocity-interpretation"><span>Calibration method helps in seismic <span class="hlt">velocity</span> interpretation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Guzman, C.E.; Davenport, H.A.; Wilhelm, R.</p> <p>1997-11-03</p> <p>Acoustic <span class="hlt">velocities</span> derived from seismic reflection data, when properly calibrated to subsurface measurements, help interpreters make pure <span class="hlt">velocity</span> predictions. A method of calibrating seismic to measured <span class="hlt">velocities</span> has improved interpretation of subsurface features in the Gulf of Mexico. In this method, the interpreter in essence creates a kind of gauge. Properly calibrated, the gauge enables the interpreter to match predicted <span class="hlt">velocities</span> to <span class="hlt">velocities</span> measured at wells. Slow-<span class="hlt">velocity</span> zones are of special interest because they sometimes appear near hydrocarbon accumulations. Changes in <span class="hlt">velocity</span> vary in strength with location; the structural picture is hidden unless the variations are accounted for by mappingmore » in depth instead of time. Preliminary <span class="hlt">observations</span> suggest that the presence of hydrocarbons alters the lithology in the neighborhood of the trap; this hydrocarbon effect may be reflected in the rock <span class="hlt">velocity</span>. The effect indicates a direct use of seismic <span class="hlt">velocity</span> in exploration. This article uses the terms seismic <span class="hlt">velocity</span> and seismic stacking <span class="hlt">velocity</span> interchangeably. It uses ground <span class="hlt">velocity</span>, checkshot average <span class="hlt">velocity</span>, and well <span class="hlt">velocity</span> interchangeably. Interval <span class="hlt">velocities</span> are derived from seismic stacking <span class="hlt">velocities</span> or well average <span class="hlt">velocities</span>; they refer to <span class="hlt">velocities</span> of subsurface intervals or zones. Interval travel time (ITT) is the reciprocal of interval <span class="hlt">velocity</span> in microseconds per foot.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AAS...22912403O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AAS...22912403O"><span>The Open Cluster Chemical Abundances and Mapping (OCCAM) Survey: Galactic Neutron Capture Abundance <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>O'Connell, Julia; Frinchaboy, Peter M.; Shetrone, Matthew D.; Melendez, Matthew; Cunha, Katia M. L.; Majewski, Steven R.; Zasowski, Gail; APOGEE Team</p> <p>2017-01-01</p> <p>The evolution of elements, as a function or age, throughout the Milky Way disk provides a key constraint for galaxy evolution models. In an effort to provide these constraints, we have conducted an investigation into the r- and s- process elemental abundances for a large sample of open clusters as part of an optical follow-up to the SDSS-III/APOGEE-1 survey. Stars were identified as cluster members by the Open Cluster Chemical Abundance & Mapping (OCCAM) survey, which culls member candidates by radial <span class="hlt">velocity</span>, metallicity, and proper motion from the <span class="hlt">observed</span> APOGEE sample. To obtain data for neutron capture elements in these clusters, we conducted a long-term <span class="hlt">observing</span> campaign covering three years (2013-2016) using the McDonald Observatory Otto Struve 2.1-m telescope and Sandiford Cass Echelle Spectrograph (R ~ 60,000). We present Galactic neutron-capture abundance <span class="hlt">gradients</span> using 30+ clusters, within 6 kpc of the Sun, covering a range of ages from ~80 Myr to ~10 Gyr .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AAS...23030706O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AAS...23030706O"><span>The Open Cluster Chemical Abundances and Mapping (OCCAM) Survey: Galactic Neutron CaptureAbundance <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>O'Connell, Julia; Frinchaboy, Peter M.; Shetrone, Matthew D.; Melendez, Matthew; Cunha, Katia; Majewski, Steven R.; Zasowski, Gail; APOGEE Team</p> <p>2017-06-01</p> <p>The evolution of elements, as a function or age, throughout the Milky Way disk provides a key constraint for galaxy evolution models. In an effort to provide these constraints, we have conducted an investigation into the r- and s- process elemental abundances for a large sample of open clusters as part of an optical follow-up to the SDSS-III/APOGEE-1 survey. Stars were identified as cluster members by the Open Cluster Chemical Abundance & Mapping (OCCAM) survey, which culls member candidates by radial <span class="hlt">velocity</span>, metallicity and proper motion from the <span class="hlt">observed</span> APOGEE sample. To obtain data for neutron capture elements in these clusters, we conducted a long-term <span class="hlt">observing</span> campaign covering three years (2013-2016) using the McDonald Observatory Otto Struve 2.1-m telescope and Sandiford Cass Echelle Spectrograph (R ~ 60,000). We present Galactic neutron capture abundance <span class="hlt">gradients</span> using 30+ clusters, within 6 kpc of the Sun, covering a range of ages from ~80 Myr to ~10 Gyr .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeoJI.210..964G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoJI.210..964G"><span>The effect of gradational <span class="hlt">velocities</span> and anisotropy on fault-zone trapped waves</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gulley, A. K.; Eccles, J. D.; Kaipio, J. P.; Malin, P. E.</p> <p>2017-08-01</p> <p>Synthetic fault-zone trapped wave (FZTW) dispersion curves and amplitude responses for FL (Love) and FR (Rayleigh) type phases are analysed in transversely isotropic 1-D elastic models. We explore the effects of <span class="hlt">velocity</span> <span class="hlt">gradients</span>, anisotropy, source location and mechanism. These experiments suggest: (i) A smooth exponentially decaying <span class="hlt">velocity</span> model produces a significantly different dispersion curve to that of a three-layer model, with the main difference being that Airy phases are not produced. (ii) The FZTW dispersion and amplitude information of a waveguide with transverse-isotropy depends mostly on the Shear wave <span class="hlt">velocities</span> in the direction parallel with the fault, particularly if the fault zone to country-rock <span class="hlt">velocity</span> contrast is small. In this low <span class="hlt">velocity</span> contrast situation, fully isotropic approximations to a transversely isotropic <span class="hlt">velocity</span> model can be made. (iii) Fault-aligned fractures and/or bedding in the fault zone that cause transverse-isotropy enhance the amplitude and wave-train length of the FR type FZTW. (iv) Moving the source and/or receiver away from the fault zone removes the higher frequencies first, similar to attenuation. (v) In most physically realistic cases, the radial component of the FR type FZTW is significantly smaller in amplitude than the transverse.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..DFDG26007M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..DFDG26007M"><span>Numerical simulation of adverse-pressure-<span class="hlt">gradient</span> boundary layer with or without roughness</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mottaghian, Pouya; Yuan, Junlin; Piomelli, Ugo</p> <p>2014-11-01</p> <p>Large-eddy and direct numerical simulations are carried out on flat-plate boundary layer over smooth and rough surfaces, with adverse pressure <span class="hlt">gradient</span>.The deceleration is achieved by imposing a wall-normal freestream <span class="hlt">velocity</span> profile, and is strong enough to cause separation at the wall. The Reynolds number based on momentum thickness and freestream <span class="hlt">velocity</span> at inlet is 600. Numerical sandgrain roughness is applied based on an immersed boundary method, yielding a flow that is transitionally rough. The turbulence intensity increases before separation, and reaches a higher value for the rough case, indicating stronger mixing. Roughness also causes higher momentum deficit near the wall, leading to earlier separation. This is consistent with previous <span class="hlt">observation</span> made on rough-wall flow separation over a ramp. In both cases, the turbulent kinetic energy peaks inside the shear layer above the detachment region, with higher values in the rough case; it then decreases approaching the reattachment region. Near the wall inside the separation bubble, the near-zero turbulent intensity indicates that the turbulent structures are lifted up in the separation region. Compared with the smooth case, the shear layer is farther from the wall and the reattachment length is longer on the rough wall.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvE..93b2305B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvE..93b2305B"><span><span class="hlt">Velocity</span> statistics of the Nagel-Schreckenberg model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bain, Nicolas; Emig, Thorsten; Ulm, Franz-Josef; Schreckenberg, Michael</p> <p>2016-02-01</p> <p>The statistics of <span class="hlt">velocities</span> in the cellular automaton model of Nagel and Schreckenberg for traffic are studied. From numerical simulations, we obtain the probability distribution function (PDF) for vehicle <span class="hlt">velocities</span> and the <span class="hlt">velocity-velocity</span> (vv) covariance function. We identify the probability to find a standing vehicle as a potential order parameter that signals nicely the transition between free congested flow for a sufficiently large number of <span class="hlt">velocity</span> states. Our results for the vv covariance function resemble features of a second-order phase transition. We develop a 3-body approximation that allows us to relate the PDFs for <span class="hlt">velocities</span> and headways. Using this relation, an approximation to the <span class="hlt">velocity</span> PDF is obtained from the headway PDF <span class="hlt">observed</span> in simulations. We find a remarkable agreement between this approximation and the <span class="hlt">velocity</span> PDF obtained from simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26986350','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26986350"><span><span class="hlt">Velocity</span> statistics of the Nagel-Schreckenberg model.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bain, Nicolas; Emig, Thorsten; Ulm, Franz-Josef; Schreckenberg, Michael</p> <p>2016-02-01</p> <p>The statistics of <span class="hlt">velocities</span> in the cellular automaton model of Nagel and Schreckenberg for traffic are studied. From numerical simulations, we obtain the probability distribution function (PDF) for vehicle <span class="hlt">velocities</span> and the <span class="hlt">velocity-velocity</span> (vv) covariance function. We identify the probability to find a standing vehicle as a potential order parameter that signals nicely the transition between free congested flow for a sufficiently large number of <span class="hlt">velocity</span> states. Our results for the vv covariance function resemble features of a second-order phase transition. We develop a 3-body approximation that allows us to relate the PDFs for <span class="hlt">velocities</span> and headways. Using this relation, an approximation to the <span class="hlt">velocity</span> PDF is obtained from the headway PDF <span class="hlt">observed</span> in simulations. We find a remarkable agreement between this approximation and the <span class="hlt">velocity</span> PDF obtained from simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1918728C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1918728C"><span>Experimental study of the free surface <span class="hlt">velocity</span> field in an asymmetrical confluence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Creelle, Stephan; Mignot, Emmanuel; Schindfessel, Laurent; De Mulder, Tom</p> <p>2017-04-01</p> <p>The hydrodynamic behavior of open channel confluences is highly complex because of the combination of different processes that interact with each other. To gain further insights in how the <span class="hlt">velocity</span> uniformization between the upstream channels and the downstream channel is proceeding, experiments are performed in a large scale 90 degree angled concrete confluence flume with a chamfered rectangular cross-section and a width of 0.98m. The dimensions and lay-out of the flume are representative for a prototype scale confluence in e.g. drainage and irrigation systems. In this type of engineered channels with sharp corners the separation zone is very large and thus the <span class="hlt">velocity</span> difference between the most contracted section and the separation zone is pronounced. With the help of surface particle tracking velocimetry the <span class="hlt">velocity</span> field is recorded from upstream of the confluence to a significant distance downstream of the confluence. The resulting data allow to analyze the evolution of the incoming flows (with a developed <span class="hlt">velocity</span> profile) that interact with the stagnation zone and each other, causing a shear layer between the two bulk flows. Close <span class="hlt">observation</span> of the <span class="hlt">velocity</span> field near the stagnation zone shows that there are actually two shear layers in the vicinity of the upstream corner. Furthermore, the data reveals that the shear layer <span class="hlt">observed</span> more downstream between the two incoming flows is actually one of the two shear layers next to the stagnation zone that continues, while the other shear layer ceases to exist. The extensive measurement domain also allows to study the shear layer between the contracted section and the separation zone. The shear layers of the stagnation zone between the incoming flows and the one between the contracted flow and separation zone are localized and parameters such as the maximum <span class="hlt">gradient</span>, <span class="hlt">velocity</span> difference and width of the shear layer are calculated. Analysis of these data shows that the shear layer between the incoming flows</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhDT.......237S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhDT.......237S"><span>Attenuation and <span class="hlt">velocity</span> dispersion in the exploration seismic frequency band</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Langqiu</p> <p></p> <p>In an anelastic medium, seismic waves are distorted by attenuation and <span class="hlt">velocity</span> dispersion, which depend on petrophysical properties of reservoir rocks. The effective attenuation and <span class="hlt">velocity</span> dispersion is a combination of intrinsic attenuation and apparent attenuation due to scattering, transmission response, and data acquisition system. <span class="hlt">Velocity</span> dispersion is usually neglected in seismic data processing partly because of insufficient <span class="hlt">observations</span> in the exploration seismic frequency band. This thesis investigates the methods of measuring <span class="hlt">velocity</span> dispersion in the exploration seismic frequency band and interprets the <span class="hlt">velocity</span> dispersion data in terms of petrophysical properties. Broadband, uncorrelated vibrator data are suitable for measuring <span class="hlt">velocity</span> dispersion in the exploration seismic frequency band, and a broad bandwidth optimizes the <span class="hlt">observability</span> of <span class="hlt">velocity</span> dispersion. Four methods of measuring <span class="hlt">velocity</span> dispersion in uncorrelated vibrator VSP data are investigated, which are the sliding window crosscorrelation (SWCC) method, the instantaneous phase method, the spectral decomposition method, and the cross spectrum method. Among them, the SWCC method is a new method and has satisfactory robustness, accuracy, and efficiency. Using the SWCC method, <span class="hlt">velocity</span> dispersion is measured in the uncorrelated vibrator VSP data from three areas with different geological settings, i.e., Mallik gas hydrate zone, McArthur River uranium mines, and Outokumpu crystalline rocks. The <span class="hlt">observed</span> <span class="hlt">velocity</span> dispersion is fitted to a straight line with respect to log frequency for a constant (frequency-independent) Q value. This provides an alternative method for calculating Q. A constant Q value does not directly link to petrophysical properties. A modeling study is implemented for the Mallik and McArthur River data to interpret the <span class="hlt">velocity</span> dispersion <span class="hlt">observations</span> in terms of petrophysical properties. The detailed multi-parameter petrophysical reservoir models are built according to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20050215631&hterms=Cluster+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DCluster%2Banalysis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050215631&hterms=Cluster+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DCluster%2Banalysis"><span>Analysis of plasmaspheric plumes: CLUSTER and IMAGE <span class="hlt">observations</span> and numerical simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Darouzet, Fabien; DeKeyser, Johan; Decreau, Pierrette; Gallagher, Dennis; Pierrard, Viviane; Lemaire, Joseph; Dandouras, Iannis; Matsui, Hiroshi; Dunlop, Malcolm; Andre, Mats</p> <p>2005-01-01</p> <p>Plasmaspheric plumes have been routinely <span class="hlt">observed</span> by CLUSTER and IMAGE. The CLUSTER mission provides high time resolution four-point measurements of the plasmasphere near perigee. Total electron density profiles can be derived from the plasma frequency and/or from the spacecraft potential (note that the electron spectrometer is usually not operating inside the plasmasphere); ion <span class="hlt">velocity</span> is also measured onboard these satellites (but ion density is not reliable because of instrumental limitations). The EUV imager onboard the IMAGE spacecraft provides global images of the plasmasphere with a spatial resolution of 0.1 RE every 10 minutes; such images acquired near apogee from high above the pole show the geometry of plasmaspheric plumes, their evolution and motion. We present coordinated <span class="hlt">observations</span> for 3 plume events and compare CLUSTER in-situ data (panel A) with global images of the plasmasphere obtained from IMAGE (panel B), and with numerical simulations for the formation of plumes based on a model that includes the interchange instability mechanism (panel C). In particular, we study the geometry and the orientation of plasmaspheric plumes by using a four-point analysis method, the spatial <span class="hlt">gradient</span>. We also compare several aspects of their motion as determined by different methods: (i) inner and outer plume boundary <span class="hlt">velocity</span> calculated from time delays of this boundary <span class="hlt">observed</span> by the wave experiment WHISPER on the four spacecraft, (ii) ion <span class="hlt">velocity</span> derived from the ion spectrometer CIS onboard CLUSTER, (iii) drift <span class="hlt">velocity</span> measured by the electron drift instrument ED1 onboard CLUSTER and (iv) global <span class="hlt">velocity</span> determined from successive EUV images. These different techniques consistently indicate that plasmaspheric plumes rotate around the Earth, with their foot fully co-rotating, but with their tip rotating slower and moving farther out.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.476.1765L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.476.1765L"><span>SDSS-IV MaNGA: global stellar population and <span class="hlt">gradients</span> for about 2000 early-type and spiral galaxies on the mass-size plane</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Hongyu; Mao, Shude; Cappellari, Michele; Ge, Junqiang; Long, R. J.; Li, Ran; Mo, H. J.; Li, Cheng; Zheng, Zheng; Bundy, Kevin; Thomas, Daniel; Brownstein, Joel R.; Roman Lopes, Alexandre; Law, David R.; Drory, Niv</p> <p>2018-05-01</p> <p>We perform full spectrum fitting stellar population analysis and Jeans Anisotropic modelling of the stellar kinematics for about 2000 early-type galaxies (ETGs) and spiral galaxies from the MaNGA DR14 sample. Galaxies with different morphologies are found to be located on a remarkably tight mass plane which is close to the prediction of the virial theorem, extending previous results for ETGs. By examining an inclined projection (`the mass-size' plane), we find that spiral and early-type galaxies occupy different regions on the plane, and their stellar population properties (i.e. age, metallicity, and stellar mass-to-light ratio) vary systematically along roughly the direction of <span class="hlt">velocity</span> dispersion, which is a proxy for the bulge fraction. Galaxies with higher <span class="hlt">velocity</span> dispersions have typically older ages, larger stellar mass-to-light ratios and are more metal rich, which indicates that galaxies increase their bulge fractions as their stellar populations age and become enriched chemically. The age and stellar mass-to-light ratio <span class="hlt">gradients</span> for low-mass galaxies in our sample tend to be positive (centre < outer), while the <span class="hlt">gradients</span> for most massive galaxies are negative. The metallicity <span class="hlt">gradients</span> show a clear peak around <span class="hlt">velocity</span> dispersion log10 σe ≈ 2.0, which corresponds to the critical mass ˜3 × 1010 M⊙ of the break in the mass-size relation. Spiral galaxies with large mass and size have the steepest <span class="hlt">gradients</span>, while the most massive ETGs, especially above the critical mass Mcrit ≳ 2 × 1011 M⊙, where slow rotator ETGs start dominating, have much flatter <span class="hlt">gradients</span>. This may be due to differences in their evolution histories, e.g. mergers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Tectp.733..148C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Tectp.733..148C"><span>Global catalog of earthquake rupture <span class="hlt">velocities</span> shows anticorrelation between stress drop and rupture <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chounet, Agnès; Vallée, Martin; Causse, Mathieu; Courboulex, Françoise</p> <p>2018-05-01</p> <p>Application of the SCARDEC method provides the apparent source time functions together with seismic moment, depth, and focal mechanism, for most of the recent earthquakes with magnitude larger than 5.6-6. Using this large dataset, we have developed a method to systematically invert for the rupture direction and average rupture <span class="hlt">velocity</span> Vr, when unilateral rupture propagation dominates. The approach is applied to all the shallow (z < 120 km) earthquakes of the catalog over the 1992-2015 time period. After a careful validation process, rupture properties for a catalog of 96 earthquakes are obtained. The subsequent analysis of this catalog provides several insights about the seismic rupture process. We first report that up-dip ruptures are more abundant than down-dip ruptures for shallow subduction interface earthquakes, which can be understood as a consequence of the material contrast between the slab and the overriding crust. Rupture <span class="hlt">velocities</span>, which are searched without any a-priori up to the maximal P wave <span class="hlt">velocity</span> (6000-8000 m/s), are found between 1200 m/s and 4500 m/s. This <span class="hlt">observation</span> indicates that no earthquakes propagate over long distances with rupture <span class="hlt">velocity</span> approaching the P wave <span class="hlt">velocity</span>. Among the 23 ruptures faster than 3100 m/s, we <span class="hlt">observe</span> both documented supershear ruptures (e.g. the 2001 Kunlun earthquake), and undocumented ruptures that very likely include a supershear phase. We also find that the correlation of Vr with the source duration scaled to the seismic moment (Ts) is very weak. This directly implies that both Ts and Vr are anticorrelated with the stress drop Δσ. This result has implications for the assessment of the peak ground acceleration (PGA) variability. As shown by Causse and Song (2015), an anticorrelation between Δσ and Vr significantly reduces the predicted PGA variability, and brings it closer to the <span class="hlt">observed</span> variability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22261656-manufacture-gradient-micro-structures-magnesium-alloys-using-two-stage-extrusion-dies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22261656-manufacture-gradient-micro-structures-magnesium-alloys-using-two-stage-extrusion-dies"><span>Manufacture of <span class="hlt">gradient</span> micro-structures of magnesium alloys using two stage extrusion dies</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hwang, Yeong-Maw; Huang, Tze-Hui; Alexandrov, Sergei</p> <p>2013-12-16</p> <p>This paper aims to manufacture magnesium alloy metals with <span class="hlt">gradient</span> micro-structures using hot extrusion process. The extrusion die was designed to have a straight channel part combined with a conical part. Materials pushed through this specially-designed die generate a non-uniform <span class="hlt">velocity</span> distribution at cross sections inside the die and result in different strain and strain rate distributions. Accordingly, a <span class="hlt">gradient</span> microstructure product can be obtained. Using the finite element analysis, the forming temperature, effective strain, and effective strain rate distributions at the die exit were firstly discussed for various inclination angles in the conical die. Then, hot extrusion experiments withmore » a two stage die were conducted to obtain magnesium alloy products with <span class="hlt">gradient</span> micro-structures. The effects of the inclination angle on the grain size distribution at cross sections of the products were also discussed. Using a die of an inclination angle of 15°, <span class="hlt">gradient</span> micro-structures of the grain size decreasing gradually from 17 μm at the center to 4 μm at the edge of product were achieved.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008MPLA...23..305L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008MPLA...23..305L"><span>Black String and <span class="hlt">Velocity</span> Frame Dragging</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Jungjai; Kim, Hyeong-Chan</p> <p></p> <p>We investigate <span class="hlt">velocity</span> frame dragging with the boosted Schwarzschild black string solution and the boosted Kaluza-Klein bubble solution, in which a translational symmetry along the boosted z-coordinate is implemented. The <span class="hlt">velocity</span> frame dragging effect can be nullified by the motion of an <span class="hlt">observer</span> using the boost symmetry along the z-coordinate if it is not compact. However, in spacetime with the compact z-coordinate, we show that the effect cannot be removed since the compactification breaks the global Lorentz boost symmetry. As a result, the comoving <span class="hlt">velocity</span> depends on r and the momentum parameter along the z-coordinate becomes an <span class="hlt">observer</span> independent characteristic quantity of the black string and bubble solutions. The dragging induces a spherical ergo-region around the black string.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/40472','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/40472"><span>Elevational <span class="hlt">gradient</span> in the cyclicity of a forest-defoliating insect</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Kyle J. Haynes; Andrew M. Liebhold; Derek M. Johnson</p> <p>2012-01-01</p> <p><span class="hlt">Observed</span> changes in the cyclicity of herbivore populations along latitudinal <span class="hlt">gradients</span> and the hypothesis that shifts in the importance of generalist versus specialist predators explain such <span class="hlt">gradients</span> has long been a matter of intense interest. In contrast, elevational <span class="hlt">gradients</span> in population cyclicity are largely unexplored. We quantified the cyclicity of gypsy moth...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100002938&hterms=car&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcar','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100002938&hterms=car&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcar"><span>High <span class="hlt">Velocity</span> Absorption during Eta Car B's Periastron Passage</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nielsen, Krister E.; Groh, J. H.; Hillier, J.; Gull, Theodore R.; Owocki, S. P.; Okazaki, A. T.; Damineli, A.; Teodoro, M.; Weigelt, G.; Hartman, H.</p> <p>2010-01-01</p> <p>Eta Car is one of the most luminous massive stars in the Galaxy, with repeated eruptions with a 5.5 year periodicity. These events are caused by the periastron passage of a massive companion in an eccentric orbit. We report the VLT/CRIRES detection of a strong high-<span class="hlt">velocity</span>, (<1900 km/s) , broad absorption wing in He I at 10833 A during the 2009.0 periastron passage. Previous <span class="hlt">observations</span> during the 2003.5 event have shown evidence of such high-<span class="hlt">velocity</span> absorption in the He I 10833 transition, allowing us to conclude that the high-<span class="hlt">velocity</span> gas is crossing the line-of-sight toward Eta Car over a time period of approximately 2 months. Our analysis of HST/STlS archival data with <span class="hlt">observations</span> of high <span class="hlt">velocity</span> absorption in the ultraviolet Si IV and C IV resonance lines, confirm the presence of a high-<span class="hlt">velocity</span> material during the spectroscopic low state. The <span class="hlt">observations</span> provide direct detection of high-<span class="hlt">velocity</span> material flowing from the wind-wind collision zone around the binary system, and we discuss the implications of the presence of high-<span class="hlt">velocity</span> gas in Eta Car during periastron</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..122.7216O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..122.7216O"><span>Sea-to-air flux of dimethyl sulfide in the South and North Pacific Ocean as measured by proton transfer reaction-mass spectrometry coupled with the <span class="hlt">gradient</span> flux technique</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Omori, Yuko; Tanimoto, Hiroshi; Inomata, Satoshi; Ikeda, Kohei; Iwata, Toru; Kameyama, Sohiko; Uematsu, Mitsuo; Gamo, Toshitaka; Ogawa, Hiroshi; Furuya, Ken</p> <p>2017-07-01</p> <p>Exchange of dimethyl sulfide (DMS) between the surface ocean and the lower atmosphere was examined by using proton transfer reaction-mass spectrometry coupled with the <span class="hlt">gradient</span> flux (PTR-MS/GF) system. We deployed the PTR-MS/GF system and <span class="hlt">observed</span> vertical <span class="hlt">gradients</span> of atmospheric DMS just above the sea surface in the subtropical and transitional South Pacific Ocean and the subarctic North Pacific Ocean. In total, we obtained 370 in situ profiles, and of these we used 46 data sets to calculate the sea-to-air flux of DMS. The DMS flux determined was in the range from 1.9 to 31 μmol m-2 d-1 and increased with wind speed and biological activity, in reasonable accordance with previous <span class="hlt">observations</span> in the open ocean. The gas transfer <span class="hlt">velocity</span> of DMS derived from the PTR-MS/GF measurements was similar to either that of DMS determined by the eddy covariance technique or that of insoluble gases derived from the dual tracer experiments, depending on the <span class="hlt">observation</span> sites located in different geographic regions. When atmospheric conditions were strongly stable during the daytime in the subtropical ocean, the PTR-MS/GF <span class="hlt">observations</span> captured a daytime versus nighttime difference in DMS mixing ratios in the surface air overlying the ocean surface. The difference was mainly due to the sea-to-air DMS emissions and stable atmospheric conditions, thus affecting the <span class="hlt">gradient</span> of DMS. This indicates that the DMS <span class="hlt">gradient</span> is strongly controlled by diurnal variations in the vertical structure of the lower atmosphere above the ocean surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010APS..MAR.K1151D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010APS..MAR.K1151D"><span><span class="hlt">Velocity</span> Dependence of the Kinetic Friction of Nanoparticles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dietzel, Dirk; Feldmann, Michael; Schirmeisen, Andre</p> <p>2010-03-01</p> <p>The <span class="hlt">velocity</span> dependence of interfacial friction is of high interest to unveil the fundamental processes in nanoscopic friction. So far, different forms of <span class="hlt">velocity</span> dependence have been <span class="hlt">observed</span> for contacts between friction force microscope (FFM) tips and a substrate surface. In this work we present <span class="hlt">velocity</span>-dependent friction measurements performed by nanoparticle manipulation of antimony nanoparticles on atomically flat HOPG substrates under UHV conditions. This allows to analyze interfacial friction for very well defined and clean surface contacts. A novel approach to nanoparticle manipulation, the so called 'tip-on-top' technique [1], made it possible to manipulate the same particle many times while varying the <span class="hlt">velocity</span>. The antimony particles exhibit a qualitatively different <span class="hlt">velocity</span> dependence on friction in comparison to direct tip-HOPG contacts. A characteristic change in <span class="hlt">velocity</span> dependence was <span class="hlt">observed</span> when comparing freshly prepared particles to contaminated specimen, which were exposed to air before the manipulation experiments. [1] Dietzel et al., Appl. Phys. Lett. 95, 53104 (2009)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870031859&hterms=Armandroff&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DArmandroff','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870031859&hterms=Armandroff&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DArmandroff"><span>The radial <span class="hlt">velocity</span>, <span class="hlt">velocity</span> dispersion, and mass-to-light ratio of the Sculptor dwarf galaxy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Armandroff, T. E.; Da Costa, G. S.</p> <p>1986-01-01</p> <p>The radial <span class="hlt">velocity</span>, <span class="hlt">velocity</span> dispersion, and mass-to-light ratio for 16 K giants in the Sculptor dwarf galaxy are calculated. Spectra at the Ca II triplet are analyzed using cross-correlation techniques in order to obtain the mean <span class="hlt">velocity</span> of + 107.4 + or - 2.0 km/s. The dimensional <span class="hlt">velocity</span> dispersion estimated as 6.3 (+1.1, -1.3) km/s is combined with the calculated core radius and <span class="hlt">observed</span> central surface brightness to produce a mass-to-light ratio of 6.0 in solar units. It is noted that the data indicate that the Sculptor contains a large amount of mass not found in globular clusters, and the mass is either in the form of remnant stars or low-mass dwarfs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMOS13D1241S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMOS13D1241S"><span>Rip Current <span class="hlt">Velocity</span> Structure in Drifter Trajectories and Numerical Simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmidt, W. E.; Slinn, D. N.</p> <p>2008-12-01</p> <p>Estimates of rip current <span class="hlt">velocity</span> and cross-shore structure were made using surfzone drifters, bathymetric surveys, and rectified video images. Over 60 rip current trajectories were <span class="hlt">observed</span> during a three year period at a Southern California beach in July 2000, 2001, and 2002. Incident wave heights (Hs) immediately offshore (~7 m depth) were obtained by initializing a refraction model with data from nearby directional wave buoys, and varied from 0.3 to 1.0 m. Tide levels varied over approximately 1 m and winds were light. Numerical simulations using the non-linear shallow water equations and modeled over measured bathymetry also produced similar flows and statistics. Time series of drifter position, sampled at 1 Hz, were first-differenced to produce <span class="hlt">velocity</span> time series. Maximum <span class="hlt">observed</span> <span class="hlt">velocities</span> varied between 25 and 80 cm s-1, whereas model maximum <span class="hlt">velocities</span> were lower by a factor 2 to 3. When <span class="hlt">velocity</span> maxima were non-dimensionalized by respective trajectory mean <span class="hlt">velocity</span>, both <span class="hlt">observed</span> and modeled values varied between 1.5 and 3.5. Cross-shore location of rip current <span class="hlt">velocity</span> maxima for both shore-normal and shore-oblique rip currents were strongly coincident with the surfzone edge (Xb), as determined by rectified video (<span class="hlt">observations</span>) or breakpoint (model). Once outside of the surfzone, <span class="hlt">observed</span> and modeled rip current <span class="hlt">velocities</span> decreased to 10% of their peak values within 2 surfzone widths of the shoreline, a useful definition of rip current cross-shore extent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ApJ...742...16T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApJ...742...16T"><span>Galaxies in ΛCDM with Halo Abundance Matching: Luminosity-<span class="hlt">Velocity</span> Relation, Baryonic Mass-<span class="hlt">Velocity</span> Relation, <span class="hlt">Velocity</span> Function, and Clustering</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Trujillo-Gomez, Sebastian; Klypin, Anatoly; Primack, Joel; Romanowsky, Aaron J.</p> <p>2011-11-01</p> <p>It has long been regarded as difficult if not impossible for a cosmological model to account simultaneously for the galaxy luminosity, mass, and <span class="hlt">velocity</span> distributions. We revisit this issue using a modern compilation of <span class="hlt">observational</span> data along with the best available large-scale cosmological simulation of dark matter (DM). We find that the standard cosmological model, used in conjunction with halo abundance matching (HAM) and simple dynamical corrections, fits—at least on average—all basic statistics of galaxies with circular <span class="hlt">velocities</span> V circ > 80 km s-1 calculated at a radius of ~10 kpc. Our primary <span class="hlt">observational</span> constraint is the luminosity-<span class="hlt">velocity</span> (LV) relation—which generalizes the Tully-Fisher and Faber-Jackson relations in allowing all types of galaxies to be included, and provides a fundamental benchmark to be reproduced by any theory of galaxy formation. We have compiled data for a variety of galaxies ranging from dwarf irregulars to giant ellipticals. The data present a clear monotonic LV relation from ~50 km s-1 to ~500 km s-1, with a bend below ~80 km s-1 and a systematic offset between late- and early-type galaxies. For comparison to theory, we employ our new ΛCDM "Bolshoi" simulation of DM, which has unprecedented mass and force resolution over a large cosmological volume, while using an up-to-date set of cosmological parameters. We use HAM to assign rank-ordered galaxy luminosities to the DM halos, a procedure that automatically fits the empirical luminosity function and provides a predicted LV relation that can be checked against <span class="hlt">observations</span>. The adiabatic contraction of DM halos in response to the infall of the baryons is included as an optional model ingredient. The resulting predictions for the LV relation are in excellent agreement with the available data on both early-type and late-type galaxies for the luminosity range from Mr = -14 to Mr = -22. We also compare our predictions for the "cold" baryon mass (i.e., stars and cold gas) of</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016CSR...113...10Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016CSR...113...10Z"><span><span class="hlt">Velocity</span> and sediment surge: What do we see at times of very shallow water on intertidal mudflats?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Qian; Gong, Zheng; Zhang, Changkuan; Townend, Ian; Jin, Chuang; Li, Huan</p> <p>2016-02-01</p> <p>A self-designed "bottom boundary layer hydrodynamic and suspended sediment concentration (SSC) measuring system" was built to <span class="hlt">observe</span> the hydrodynamic and the SSC processes over the intertidal mudflats at the middle part of the Jiangsu coast during August 8-10, 2013. <span class="hlt">Velocity</span> profiles within 10 cm of the mudflat surface were obtained with a vertical resolution as fine as 1 mm. An ADCP was used to extend the profile over the full water depth with a resolution of 10 cm and the vertical SSC profile was measured at intervals using Optical Backscatter Sensors (OBS). At the same time, water levels and wave conditions were measured with a Tide and Wave Recorder. Measured data suggested that the vertical structure of <span class="hlt">velocity</span> profiles within 10 cm above the bed maintains a logarithmic distribution during the whole tidal cycle except the slack-water periods. Shallow flows during both the early-flood period and the later-ebb period are characterized by a relatively large vertical <span class="hlt">velocity</span> <span class="hlt">gradient</span> and a "surge" feature. We conclude that the very shallow water stages are transient and may not contribute much to the whole water and sediment transport, while they can play a significant role in the formation and evolution of micro-topographies on tidal flats.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22718875','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22718875"><span>Effect of <span class="hlt">gradient</span> dielectric coefficient in a functionally graded material (FGM) substrate on the propagation behavior of love waves in an FGM-piezoelectric layered structure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cao, Xiaoshan; Shi, Junping; Jin, Feng</p> <p>2012-06-01</p> <p>The propagation behavior of Love waves in a layered structure that includes a functionally graded material (FGM) substrate carrying a piezoelectric thin film is investigated. Analytical solutions are obtained for both constant and <span class="hlt">gradient</span> dielectric coefficients in the FGM substrate. Numerical results show that the <span class="hlt">gradient</span> dielectric coefficient decreases phase <span class="hlt">velocity</span> in any mode, and the electromechanical coupling factor significantly increases in the first- and secondorder modes. In some modes, the difference in Love waves' phase <span class="hlt">velocity</span> between these two types of structure might be more than 1%, resulting in significant differences in frequency of the surface acoustic wave devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ISPAr42W7.1477Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ISPAr42W7.1477Z"><span>Analyzing the <span class="hlt">Velocity</span> of Urban Dynamic Over Northeastern China Using Dmsp-Ols Night-Time Lights</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Y.</p> <p>2017-09-01</p> <p>Stable night-time lights (NTL) data from the Defense Meteorological Satellite Program Operational Line-scan System (DMSPOLS) can serve as a good proxy for anthropogenic development. Here DMSP-OLS NTL data was used to detect the urban development status in northeastern China. The spatial and temporal <span class="hlt">gradients</span> are combined to depict the <span class="hlt">velocity</span> of urban expanding process. This <span class="hlt">velocity</span> index represents the instantaneous local <span class="hlt">velocity</span> along the Earth's surface needed to maintain constant NTL condition, and has a mean of 0.36 km/yr for northeastern China. The <span class="hlt">velocity</span> change of NTL is lower in the urban center and its near regions, and the suburbs show a relatively high value. The connecting zones between satellite cities and metropolis have also a rapid rate of NTL evolution. The dynamic process of urbanization over the study area is mainly in a manner of spreading from urban cores to edges. The rank size of the <span class="hlt">velocity</span> for the prefectures is analyzed and a long tail distribution is found. The <span class="hlt">velocity</span> index can provide insights for the future pattern of urban sprawl.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17678096','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17678096"><span>Toroidal momentum pinch <span class="hlt">velocity</span> due to the coriolis drift effect on small scale instabilities in a toroidal plasma.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Peeters, A G; Angioni, C; Strintzi, D</p> <p>2007-06-29</p> <p>In this Letter, the influence of the "Coriolis drift" on small scale instabilities in toroidal plasmas is shown to generate a toroidal momentum pinch <span class="hlt">velocity</span>. Such a pinch results because the Coriolis drift generates a coupling between the density and temperature perturbations on the one hand and the perturbed parallel flow <span class="hlt">velocity</span> on the other. A simple fluid model is used to highlight the physics mechanism and gyro-kinetic calculations are performed to accurately assess the magnitude of the pinch. The derived pinch <span class="hlt">velocity</span> leads to a radial <span class="hlt">gradient</span> of the toroidal <span class="hlt">velocity</span> profile even in the absence of a torque on the plasma and is predicted to generate a peaking of the toroidal <span class="hlt">velocity</span> profile similar to the peaking of the density profile. Finally, the pinch also affects the interpretation of current experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSEC21A..02R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSEC21A..02R"><span><span class="hlt">Observations</span> of Inner Shelf Flows Influenced by a Small-Scale River Plume in the Northern Gulf of Mexico</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roth, M.; MacMahan, J.; Reniers, A.; Ozgokmen, T. M.</p> <p>2016-02-01</p> <p>Recent work has demonstrated that wind and waves are important forcing mechanisms for the inner shelf vertical current structure. Here, the inner shelf flows are evaluated away from an adjacent inlet where a small-scale buoyant plume emerges. The plume's nearshore extent, speed, vertical thickness, and density are controlled by the passage of low-pressure extratropical cyclones that are common in the northern Gulf of Mexico. The colder, brackish plume water provides vertical stratification and a cross-shore density <span class="hlt">gradient</span> with the warmer, saline oceanic water. An Acoustic Doppler Current Profiler (ADCP) was deployed in 10m water depth as part of an intensive 2-week experiment (SCOPE), which also obtained wind and cross-shelf temperature, salinity, and <span class="hlt">velocity</span>. The 10m ADCP remained collecting an additional year of <span class="hlt">velocity</span> <span class="hlt">observations</span>. The plume was not always present, but episodically influenced the experiment site. When the plume reached the site, the alongshore surface and subsurface typically flowed in opposite directions, likely caused by plume-induced pressure <span class="hlt">gradients</span>. Plumes that extended into the subsurface appear to have caused depth-averaged onshore flow above that expected from wind and wave-driven forcing. <span class="hlt">Observations</span> from SCOPE and the 1-year ADCP are used to describe seasonal full-depth flow patterns influenced by wind, waves, and plume presence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930080742','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930080742"><span><span class="hlt">Observations</span> on the method of determining the <span class="hlt">velocity</span> of airships</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Volterra, Vito</p> <p>1921-01-01</p> <p>To obtain the absolute <span class="hlt">velocity</span> of an airship by knowing the speed at which two routes are covered, we have only to determine the geographical direction of the routes which we locate from a map, and the angles of routes as given by the compass, after correcting for the variation (the algebraical sum of the local magnetic declination and the deviation).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RSPSA.47470706W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RSPSA.47470706W"><span>Effects of Lewis number on the statistics of the invariants of the <span class="hlt">velocity</span> <span class="hlt">gradient</span> tensor and local flow topologies in turbulent premixed flames</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wacks, Daniel; Konstantinou, Ilias; Chakraborty, Nilanjan</p> <p>2018-04-01</p> <p>The behaviours of the three invariants of the <span class="hlt">velocity</span> <span class="hlt">gradient</span> tensor and the resultant local flow topologies in turbulent premixed flames have been analysed using three-dimensional direct numerical simulation data for different values of the characteristic Lewis number ranging from 0.34 to 1.2. The results have been analysed to reveal the statistical behaviours of the invariants and the flow topologies conditional upon the reaction progress variable. The behaviours of the invariants have been explained in terms of the relative strengths of the thermal and mass diffusions, embodied by the influence of the Lewis number on turbulent premixed combustion. Similarly, the behaviours of the flow topologies have been explained in terms not only of the Lewis number but also of the likelihood of the occurrence of individual flow topologies in the different flame regions. Furthermore, the sensitivity of the joint probability density function of the second and third invariants and the joint probability density functions of the mean and Gaussian curvatures to the variation in Lewis number have similarly been examined. Finally, the dependences of the scalar-turbulence interaction term on augmented heat release and of the vortex-stretching term on flame-induced turbulence have been explained in terms of the Lewis number, flow topology and reaction progress variable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5938671','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5938671"><span>Effects of Lewis number on the statistics of the invariants of the <span class="hlt">velocity</span> <span class="hlt">gradient</span> tensor and local flow topologies in turbulent premixed flames</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Konstantinou, Ilias; Chakraborty, Nilanjan</p> <p>2018-01-01</p> <p>The behaviours of the three invariants of the <span class="hlt">velocity</span> <span class="hlt">gradient</span> tensor and the resultant local flow topologies in turbulent premixed flames have been analysed using three-dimensional direct numerical simulation data for different values of the characteristic Lewis number ranging from 0.34 to 1.2. The results have been analysed to reveal the statistical behaviours of the invariants and the flow topologies conditional upon the reaction progress variable. The behaviours of the invariants have been explained in terms of the relative strengths of the thermal and mass diffusions, embodied by the influence of the Lewis number on turbulent premixed combustion. Similarly, the behaviours of the flow topologies have been explained in terms not only of the Lewis number but also of the likelihood of the occurrence of individual flow topologies in the different flame regions. Furthermore, the sensitivity of the joint probability density function of the second and third invariants and the joint probability density functions of the mean and Gaussian curvatures to the variation in Lewis number have similarly been examined. Finally, the dependences of the scalar--turbulence interaction term on augmented heat release and of the vortex-stretching term on flame-induced turbulence have been explained in terms of the Lewis number, flow topology and reaction progress variable. PMID:29740257</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AMT.....9.2581R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AMT.....9.2581R"><span>Application of an online ion-chromatography-based instrument for <span class="hlt">gradient</span> flux measurements of speciated nitrogen and sulfur</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rumsey, Ian C.; Walker, John T.</p> <p>2016-06-01</p> <p>The dry component of total nitrogen and sulfur atmospheric deposition remains uncertain. The lack of measurements of sufficient chemical speciation and temporal extent make it difficult to develop accurate mass budgets and sufficient process level detail is not available to improve current air-surface exchange models. Over the past decade, significant advances have been made in the development of continuous air sampling measurement techniques, resulting with instruments of sufficient sensitivity and temporal resolution to directly quantify air-surface exchange of nitrogen and sulfur compounds. However, their applicability is generally restricted to only one or a few of the compounds within the deposition budget. Here, the performance of the Monitor for AeRosols and GAses in ambient air (MARGA 2S), a commercially available online ion-chromatography-based analyzer is characterized for the first time as applied for air-surface exchange measurements of HNO3, NH3, NH4+, NO3-, SO2 and SO42-. Analytical accuracy and precision are assessed under field conditions. Chemical concentrations <span class="hlt">gradient</span> precision are determined at the same sampling site. Flux uncertainty measured by the aerodynamic <span class="hlt">gradient</span> method is determined for a representative 3-week period in fall 2012 over a grass field. Analytical precision and chemical concentration <span class="hlt">gradient</span> precision were found to compare favorably in comparison to previous studies. During the 3-week period, percentages of hourly chemical concentration <span class="hlt">gradients</span> greater than the corresponding chemical concentration <span class="hlt">gradient</span> detection limit were 86, 42, 82, 73, 74 and 69 % for NH3, NH4+, HNO3, NO3-, SO2 and SO42-, respectively. As expected, percentages were lowest for aerosol species, owing to their relatively low deposition <span class="hlt">velocities</span> and correspondingly smaller <span class="hlt">gradients</span> relative to gas phase species. Relative hourly median flux uncertainties were 31, 121, 42, 43, 67 and 56 % for NH3, NH4+, HNO3, NO3-, SO2 and SO42-, respectively. Flux</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25424492','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25424492"><span>Near field plasmonic <span class="hlt">gradient</span> effects on high vacuum tip-enhanced Raman spectroscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fang, Yurui; Zhang, Zhenglong; Chen, Li; Sun, Mengtao</p> <p>2015-01-14</p> <p>Near field <span class="hlt">gradient</span> effects in high vacuum tip-enhanced Raman spectroscopy (HV-TERS) are a recent developing ultra-sensitive optical and spectral analysis technology on the nanoscale, based on the plasmons and plasmonic <span class="hlt">gradient</span> enhancement in the near field and under high vacuum. HV-TERS can not only be used to detect ultra-sensitive Raman spectra enhanced by surface plasmon, but also to detect clear molecular IR-active modes enhanced by strongly plasmonic <span class="hlt">gradient</span>. Furthermore, the molecular overtone modes and combinational modes can also be experimentally measured, where the Fermi resonance and Darling-Dennison resonance were successfully <span class="hlt">observed</span> in HV-TERS. Theoretical calculations using electromagnetic field theory firmly supported experimental <span class="hlt">observation</span>. The intensity ratio of the plasmon <span class="hlt">gradient</span> term over the linear plasmon term can reach values greater than 1. Theoretical calculations also revealed that with the increase in gap distance between tip and substrate, the decrease in the plasmon <span class="hlt">gradient</span> was more significant than the decrease in plasmon intensity, which is the reason that the <span class="hlt">gradient</span> Raman can be only <span class="hlt">observed</span> in the near field. Recent experimental results of near field <span class="hlt">gradient</span> effects on HV-TERS were summarized, following the section of the theoretical analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDH35001C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDH35001C"><span>A dissipative random <span class="hlt">velocity</span> field for fully developed fluid turbulence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chevillard, Laurent; Pereira, Rodrigo; Garban, Christophe</p> <p>2016-11-01</p> <p>We investigate the statistical properties, based on numerical simulations and analytical calculations, of a recently proposed stochastic model for the <span class="hlt">velocity</span> field of an incompressible, homogeneous, isotropic and fully developed turbulent flow. A key step in the construction of this model is the introduction of some aspects of the vorticity stretching mechanism that governs the dynamics of fluid particles along their trajectory. An additional further phenomenological step aimed at including the long range correlated nature of turbulence makes this model depending on a single free parameter that can be estimated from experimental measurements. We confirm the realism of the model regarding the geometry of the <span class="hlt">velocity</span> <span class="hlt">gradient</span> tensor, the power-law behaviour of the moments of <span class="hlt">velocity</span> increments, including the intermittent corrections, and the existence of energy transfers across scales. We quantify the dependence of these basic properties of turbulent flows on the free parameter and derive analytically the spectrum of exponents of the structure functions in a simplified non dissipative case. A perturbative expansion shows that energy transfers indeed take place, justifying the dissipative nature of this random field.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20432304','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20432304"><span>Steering of aggregating magnetic microparticles using propulsion <span class="hlt">gradients</span> coils in an MRI Scanner.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mathieu, Jean-Baptiste; Martel, Sylvain</p> <p>2010-05-01</p> <p>Upgraded <span class="hlt">gradient</span> coils can effectively enhance the MRI steering of magnetic microparticles in a branching channel. Applications of this method include MRI targeting of magnetic embolization agents for oncologic therapy. A magnetic suspension of Fe(3)O(4) magnetic particles was injected inside a y-shaped microfluidic channel. Magnetic <span class="hlt">gradients</span> of 0, 50, 100, 200, and 400 mT/m were applied to the magnetic particles perpendicularly to the flow by a custom-built <span class="hlt">gradient</span> coil inside a 1.5-T MRI scanner. Measurement of the steering ratio was performed both by video analyses and quantification of the mass of the particles collected at each outlet of the microfluidic channel, using atomic absorption spectroscopy. Magnetic particles steering ratios of 0.99 and 0.75 were reached with 400 mT/m <span class="hlt">gradient</span> amplitude and measured by video analyses and atomic absorption spectroscopy, respectively. Experimental data shows that the steering ratio increases with higher magnetic <span class="hlt">gradients</span>. Moreover, theory suggests that larger particles (or aggregates), higher magnetizations, and lower flows can also be used to improve the steering ratio. The technological limitation of the approach is that an MRI <span class="hlt">gradient</span> amplitude increase to a few hundred milliteslas per meter is needed. A simple analytical method based on magnetophoretic <span class="hlt">velocity</span> predictions and geometric considerations is proposed for steering ratio calculation. (c) 2010 Wiley-Liss, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22086180-alfven-critical-ionization-velocity-observed-high-power-impulse-magnetron-sputtering-discharges','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22086180-alfven-critical-ionization-velocity-observed-high-power-impulse-magnetron-sputtering-discharges"><span>Alfven's critical ionization <span class="hlt">velocity</span> <span class="hlt">observed</span> in high power impulse magnetron sputtering discharges</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Brenning, N.; Lundin, D.</p> <p>2012-09-15</p> <p>Azimuthally rotating dense plasma structures, spokes, have recently been detected in several high power impulse magnetron sputtering (HiPIMS) devices used for thin film deposition and surface treatment, and are thought to be important for plasma buildup, energizing of electrons, as well as cross-B transport of charged particles. In this work, the drift <span class="hlt">velocities</span> of these spokes are shown to be strongly correlated with the critical ionization <span class="hlt">velocity</span>, CIV, proposed by Alfven. It is proposed as the most promising approach in combining the CIV and HiPIMS research fields is to focus on the role of spokes in the process of electronmore » energization.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1332903-observation-acceleration-deceleration-gigaelectron-volt-per-metre-gradient-dielectric-wakefield-accelerators','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1332903-observation-acceleration-deceleration-gigaelectron-volt-per-metre-gradient-dielectric-wakefield-accelerators"><span><span class="hlt">Observation</span> of acceleration and deceleration in gigaelectron-volt-per-metre <span class="hlt">gradient</span> dielectric wakefield accelerators</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>O’Shea, B. D.; Andonian, G.; Barber, S. K.; ...</p> <p>2016-09-14</p> <p>There is urgent need to develop new acceleration techniques capable of exceeding gigaelectron-volt-per-metre (GeV m –1) <span class="hlt">gradients</span> in order to enable future generations of both light sources and high-energy physics experiments. To address this need, short wavelength accelerators based on wakefields, where an intense relativistic electron beam radiates the demanded fields directly into the accelerator structure or medium, are currently under intense investigation. One such wakefield based accelerator, the dielectric wakefield accelerator, uses a dielectric lined-waveguide to support a wakefield used for acceleration. Here we show <span class="hlt">gradients</span> of 1.347±0.020 GeV m –1 using a dielectric wakefield accelerator of 15 cmmore » length, with sub-millimetre transverse aperture, by measuring changes of the kinetic state of relativistic electron beams. We follow this measurement by demonstrating accelerating <span class="hlt">gradients</span> of 320±17 MeV m –1. As a result, both measurements improve on previous measurements by and order of magnitude and show promise for dielectric wakefield accelerators as sources of high-energy electrons.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5027279','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5027279"><span><span class="hlt">Observation</span> of acceleration and deceleration in gigaelectron-volt-per-metre <span class="hlt">gradient</span> dielectric wakefield accelerators</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>O'Shea, B. D.; Andonian, G.; Barber, S. K.; Fitzmorris, K. L.; Hakimi, S.; Harrison, J.; Hoang, P. D.; Hogan, M. J.; Naranjo, B.; Williams, O. B.; Yakimenko, V.; Rosenzweig, J. B.</p> <p>2016-01-01</p> <p>There is urgent need to develop new acceleration techniques capable of exceeding gigaelectron-volt-per-metre (GeV m−1) <span class="hlt">gradients</span> in order to enable future generations of both light sources and high-energy physics experiments. To address this need, short wavelength accelerators based on wakefields, where an intense relativistic electron beam radiates the demanded fields directly into the accelerator structure or medium, are currently under intense investigation. One such wakefield based accelerator, the dielectric wakefield accelerator, uses a dielectric lined-waveguide to support a wakefield used for acceleration. Here we show <span class="hlt">gradients</span> of 1.347±0.020 GeV m−1 using a dielectric wakefield accelerator of 15 cm length, with sub-millimetre transverse aperture, by measuring changes of the kinetic state of relativistic electron beams. We follow this measurement by demonstrating accelerating <span class="hlt">gradients</span> of 320±17 MeV m−1. Both measurements improve on previous measurements by and order of magnitude and show promise for dielectric wakefield accelerators as sources of high-energy electrons. PMID:27624348</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S34A..08T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S34A..08T"><span>Spatial and temporal variation of seismic <span class="hlt">velocity</span> during earthquakes and volcanic eruptions in western Japan: Insight into mechanism for seismic <span class="hlt">velocity</span> variation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsuji, T.; Ikeda, T.; Nimiya, H.</p> <p>2017-12-01</p> <p>We report spatio-temporal variations of seismic <span class="hlt">velocity</span> around the seismogenic faults in western Japan. We mainly focus on the seismic <span class="hlt">velocity</span> variation during (1) the 2016 Off-Mie earthquake in the Nankai subduction zone (Mw5.8) and (2) the 2016 Kumamoto earthquake in Kyushu Island (Mw7.0). We applied seismic interferometry and surface wave analysis to the ambient noise data recorded by Hi-net and DONET seismometers of National Research Institute for Earth Science and Disaster Resilience (NIED). Seismic <span class="hlt">velocity</span> near the rupture faults and volcano decreased during the earthquake. For example, we <span class="hlt">observed</span> <span class="hlt">velocity</span> reduction around the seismogenic Futagawa-Hinagu fault system and Mt Aso in the 2016 Kumamoto earthquake. We also identified <span class="hlt">velocity</span> increase after the eruptions of Mt Aso. During the 2016 Off-Mie earthquake, we <span class="hlt">observed</span> seismic <span class="hlt">velocity</span> variation in the Nankai accretionary prism. After the earthquakes, the seismic <span class="hlt">velocity</span> gradually returned to the pre-earthquake value. The <span class="hlt">velocity</span> recovering process (healing process) is caused by several mechanisms, such as pore pressure reduction, strain change, and crack sealing. By showing the <span class="hlt">velocity</span> variations obtained at different geologic settings (volcano, seismogenic fault, unconsolidated sediment), we discuss the mechanism of seismic <span class="hlt">velocity</span> variation as well as the post-seismic fault healing process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSM51F2566A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSM51F2566A"><span>Particle-in-cell Simulations of Waves in a Plasma Described by Kappa <span class="hlt">Velocity</span> Distribution as <span class="hlt">Observed</span> in the Saturńs Magnetosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alves, M. V.; Barbosa, M. V. G.; Simoes, F. J. L., Jr.</p> <p>2016-12-01</p> <p><span class="hlt">Observations</span> have shown that several regions in space plasmas exhibit non-Maxwellian distributions with high energy superthermal tails. Kappa <span class="hlt">velocity</span> distribution functions can describe many of these regions and have been used since the 60's. They suit well to represent superthermal tails in solar wind as well as to obtain plasma parameters of plasma within planetary magnetospheres. A set of initial <span class="hlt">velocities</span> following kappa distribution functions is used in KEMPO1 particle simulation code to analyze the normal modes of wave propagation. Initial conditions are determined using <span class="hlt">observed</span> characteristics for Saturńs magnetosphere. Two electron species with different temperatures and densities and ions as a third species are used. Each electron population is described by a different kappa index. Particular attention is given to perpendicular propagation, Bernstein modes, and parallel propagation, Langmuir and electron-acoustic modes. The dispersion relation for the Bernstein modes is strongly influenced by the shape of the <span class="hlt">velocity</span> distribution and consequently by the value of kappa index. Simulation results are compared with numerical solutions of the dispersion relation obtained in the literature and they are in good agreement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997PhDT.......219L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997PhDT.......219L"><span>Combining deterministic and stochastic <span class="hlt">velocity</span> fields in the analysis of deep crustal seismic data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Larkin, Steven Paul</p> <p></p> <p>Standard crustal seismic modeling obtains deterministic <span class="hlt">velocity</span> models which ignore the effects of wavelength-scale heterogeneity, known to exist within the Earth's crust. Stochastic <span class="hlt">velocity</span> models are a means to include wavelength-scale heterogeneity in the modeling. These models are defined by statistical parameters obtained from geologic maps of exposed crystalline rock, and are thus tied to actual geologic structures. Combining both deterministic and stochastic <span class="hlt">velocity</span> models into a single model allows a realistic full wavefield (2-D) to be computed. By comparing these simulations to recorded seismic data, the effects of wavelength-scale heterogeneity can be investigated. Combined deterministic and stochastic <span class="hlt">velocity</span> models are created for two datasets, the 1992 RISC seismic experiment in southeastern California and the 1986 PASSCAL seismic experiment in northern Nevada. The RISC experiment was located in the transition zone between the Salton Trough and the southern Basin and Range province. A high-<span class="hlt">velocity</span> body previously identified beneath the Salton Trough is constrained to pinch out beneath the Chocolate Mountains to the northeast. The lateral extent of this body is evidence for the ephemeral nature of rifting loci as a continent is initially rifted. Stochastic modeling of wavelength-scale structures above this body indicate that little more than 5% mafic intrusion into a more felsic continental crust is responsible for the <span class="hlt">observed</span> reflectivity. Modeling of the wide-angle RISC data indicates that coda waves following PmP are initially dominated by diffusion of energy out of the near-surface basin as the wavefield reverberates within this low-<span class="hlt">velocity</span> layer. At later times, this coda consists of scattered body waves and P to S conversions. Surface waves do not play a significant role in this coda. Modeling of the PASSCAL dataset indicates that a high-<span class="hlt">gradient</span> crust-mantle transition zone or a rough Moho interface is necessary to reduce precritical Pm</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDD19003D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDD19003D"><span>The relation between skin friction fluctuations and turbulent fluctuating <span class="hlt">velocities</span> in turbulent boundary layers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Diaz Daniel, Carlos; Laizet, Sylvain; Vassilicos, John Christos</p> <p>2015-11-01</p> <p>The Townsend-Perry hypothesis of wall-attached eddies relates the friction <span class="hlt">velocity</span> uτ at the wall to <span class="hlt">velocity</span> fluctuations at a position y from the wall, resulting in a wavenumber range where the streamwise fluctuating <span class="hlt">velocity</span> spectrum scales as E (k) ~k-1 and the corresponding structure function scales as uτ2 in the corresponding length-scale range. However, this model does not take in account the fluctuations of the skin friction <span class="hlt">velocity</span>, which are in fact strongly intermittent. A DNS of zero-pressure <span class="hlt">gradient</span> turbulent boundary layer suggests a 10 to 15 degree angle from the lag of the peak in the cross-correlations between the fluctuations of the shear stress and streamwise fluctuating <span class="hlt">velocities</span> at different heights in the boundary layer. Using this result, it is possible to refine the definition of the attached eddy range of scales, and our DNS suggests that, in this range, the second order structure function depends on filtered skin friction fluctuations in a way which is about the same at different distances from the wall and different local Reynolds numbers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AAS...21536304S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AAS...21536304S"><span>Demography of SDSS Early-type Galaxies from the Perspective of Radial Color <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suh, Hyewon; Jeong, H.; Oh, K.; Yi, S. K.; Ferreras, I.; Schawinski, K.</p> <p>2010-01-01</p> <p>We have investigated the radial g-r color <span class="hlt">gradients</span> of early-type galaxies in the Sloan Digital Sky Survey (SDSS) DR6 in the redshift range 0.00 < z < 0.06. The majority of massive early-type galaxies show a negative color <span class="hlt">gradient</span> (centers being redder). On the other hand, roughly 30 percent of the galaxies in this sample show positive color <span class="hlt">gradients</span> (centers being bluer). These positive-<span class="hlt">gradient</span> galaxies often show strong Hβ absorption line strengths and/or emission line ratios that are consistent with containing young stellar populations. Combining the optical data with Galaxy Evolution Explorer (GALEX) UV photometry, we find that all positive-<span class="hlt">gradient</span> galaxies show blue UV-optical colors. This implies that the residual star formation in early-type galaxies is centrally concentrated. These positive-<span class="hlt">gradient</span> galaxies tend to live in lower density regions. They are also a bit more likely to have a late-type companion galaxy, hinting at a possible role of interactions with a gas-rich companion. A simplistic population analysis shows that these positive color <span class="hlt">gradients</span> are visible only for half a billion years after a star burst. Moreover, the positive-<span class="hlt">gradient</span> galaxies occupy different regions in the fundamental planes from the outnumbering negative-<span class="hlt">gradient</span> galaxies. However, the positions of the positive-<span class="hlt">gradient</span> galaxies on the fundamental planes cannot be attributed to any reasonable amount of recent star formation alone but require substantially lower <span class="hlt">velocity</span> dispersions to begin with. Our results based on the optical data are consistent with the residual star formation interpretation which was based on the GALEX UV data. A low-level residual star formation seems continuing in most of the less-massive early-type galaxies in their centers.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JCoPh.330..650M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JCoPh.330..650M"><span>A parabolic <span class="hlt">velocity</span>-decomposition method for wind turbines</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mittal, Anshul; Briley, W. Roger; Sreenivas, Kidambi; Taylor, Lafayette K.</p> <p>2017-02-01</p> <p>An economical parabolized Navier-Stokes approximation for steady incompressible flow is combined with a compatible wind turbine model to simulate wind turbine flows, both upstream of the turbine and in downstream wake regions. The inviscid parabolizing approximation is based on a Helmholtz decomposition of the secondary <span class="hlt">velocity</span> vector and physical order-of-magnitude estimates, rather than an axial pressure <span class="hlt">gradient</span> approximation. The wind turbine is modeled by distributed source-term forces incorporating time-averaged aerodynamic forces generated by a blade-element momentum turbine model. A solution algorithm is given whose dependent variables are streamwise <span class="hlt">velocity</span>, streamwise vorticity, and pressure, with secondary <span class="hlt">velocity</span> determined by two-dimensional scalar and vector potentials. In addition to laminar and turbulent boundary-layer test cases, solutions for a streamwise vortex-convection test problem are assessed by mesh refinement and comparison with Navier-Stokes solutions using the same grid. Computed results for a single turbine and a three-turbine array are presented using the NREL offshore 5-MW baseline wind turbine. These are also compared with an unsteady Reynolds-averaged Navier-Stokes solution computed with full rotor resolution. On balance, the agreement in turbine wake predictions for these test cases is very encouraging given the substantial differences in physical modeling fidelity and computer resources required.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSMG13A..05D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSMG13A..05D"><span>Hydrodynamic Controls on Muddy Sedimentary Fabric Development on Low-<span class="hlt">Gradient</span> Shelves: Atchafalaya Chenier Plain Subaqueous Delta</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Denommee, K.; Bentley, S. J.; Harazim, D.; Macquaker, J.</p> <p>2016-02-01</p> <p>Short sediment cores and geophysical data collected on the Southwest Louisiana Chenier Plain inner shelf have been studied in order to examine the sedimentary products of current-wave-enhanced sediment gravity flows (CWESGFs), a type of sediment gravity flow where the driving energy required to transport sediment across low-<span class="hlt">gradient</span> settings is augmented by the near-bed orbital <span class="hlt">velocity</span> of surface gravity wave and near-bed currents. Sedimentary fabrics <span class="hlt">observed</span> on the SWLA shelf document the following flow evolution: (1) the erosion of the underlying substrate in response to wave-generated shear stresses in the bottom boundary layer, followed by (2) the deposition of ripple a crossbeded unit during wave-mediated oscillatory motions in low-viscosity suspension; (3) the deposition of subtle intercalated laminae during laminar flow at higher suspended sediment concentrations; followed by the deposition of (4) normally graded sediments during the waning phases of the flow. Significantly, the sedimentary fabrics deposited by CWESGFs on SWLA shelf show diagnostic variations from CWESGF-generated sedimentary fabrics <span class="hlt">observed</span> on the Eel and Amazon shelves. Differences between the <span class="hlt">observed</span> sedimentary fabrics are hypothesized to result from variations in the relative contribution of near-bed currents, wave orbital <span class="hlt">velocities</span>, and bed slope (gravity) to the driving energy of the CWESGF, and as such can be catalogued as diagnostic recognition criteria using a prismatic ternary diagram where current-, wave-, and gravity-dominated end members form the vertices of a triangle, and wave period forms the prism axis. In this framework forcing mechanisms can be represented quantitatively, based on wave period and the relative contribution of each of the CWESGF <span class="hlt">velocity</span> terms. This framework can be used to explore relationships between hydrodynamics and CWESGF fabrics, providing geologists with a tool with which to better recognize the depositional products of CWESGFs in the rock</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.T23B2587P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.T23B2587P"><span>Integrating shear <span class="hlt">velocity</span> <span class="hlt">observations</span> of the Hudson Bay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Porritt, R. W.; Miller, M. S.; Darbyshire, F. A.</p> <p>2013-12-01</p> <p>Hudson Bay is the core of the Laurentia craton of North America. This region contains some of the thickest lithosphere globally, reaching 250-300 km depth. Previous studies have shown that much of this region is composed of amalgamated proto-continents including the Western Churchill and Superior provinces and that much of the structure of these constituents has been retained since the Trans-Hudson Orogen at 1.8 Ga. Using the Hudson Bay Lithospheric Experiment (HuBLE) and other permanent and POLARIS broadband seismic data, we image the region with S to P receiver functions, joint inversion of P to S receiver functions with surface waves, and teleseismic S and P wave travel-times. The receiver function imaging reveals a persistent mid-lithospheric layer at ~80 km depth under all stations, but a variable lithospheric thickness. The teleseismic S delay times show a pattern of early arrivals around the center of the network, beneath Hudson Bay where the lithosphere is thickest, while the P delay times are early in the Superior province relative to the Western Churchill province. This suggests higher Vp/Vs ratios in the Superior province, which is evidence that stacked oceanic plates formed this province. The relatively flat Moho imaged by earlier receiver function studies and the lower mantle Vp/Vs of the Western Churchill province provides evidence of formation by plume head extraction. The joint inversion shows an LAB that is typically a broad discontinuity spanning ~20-30 km at ~220 km depth suggesting a primarily thermal boundary zone. The mid-lithospheric layer is composed of increasing <span class="hlt">velocity</span> from the ~40 km depth Moho defined by H-k stacking of PRFs to a broad, constant <span class="hlt">velocity</span> lithospheric lid spanning 80-200 km depth. We suggest this mid-lithospheric layer represents the mantle lithosphere of the proto-continents prior to collision and the lid formed due to post-collisional cooling. The integration of these seismic datasets furthers our understanding of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840005054','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840005054"><span><span class="hlt">Velocity</span> profiles of interplanetary shocks</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cane, H. V.</p> <p>1983-01-01</p> <p>The type 2 radio burst was identified as a shock propagating through solar corona. Radio emission from shocks travelling through the interplanetary (IP) medium was <span class="hlt">observed</span>. Using the drift rates of IP type II bursts the <span class="hlt">velocity</span> characteristics of eleven shocks were investigated. It is indicated that shocks in the IP medium undergo acceleration before decelerating and that the slower shocks take longer to attain their maximum <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Gradient&id=EJ827263','ERIC'); return false;" href="https://eric.ed.gov/?q=Gradient&id=EJ827263"><span>Generalizability of Scaling <span class="hlt">Gradients</span> on Direct Behavior Ratings</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Chafouleas, Sandra M.; Christ, Theodore J.; Riley-Tillman, T. Chris</p> <p>2009-01-01</p> <p>Generalizability theory is used to examine the impact of scaling <span class="hlt">gradients</span> on a single-item Direct Behavior Rating (DBR). A DBR refers to a type of rating scale used to efficiently record target behavior(s) following an <span class="hlt">observation</span> occasion. Variance components associated with scale <span class="hlt">gradients</span> are estimated using a random effects design for persons…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25998723','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25998723"><span>Neutrophil migration under spatially-varying chemoattractant <span class="hlt">gradient</span> profiles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Halilovic, Iris; Wu, Jiandong; Alexander, Murray; Lin, Francis</p> <p>2015-01-01</p> <p>Chemotaxis plays an important role in biological processes such as cancer metastasis, embryogenesis, wound healing, and immune response. Neutrophils are the frontline defenders against invasion of foreign microorganisms into our bodies. To achieve this important immune function, a neutrophil can sense minute chemoattractant concentration differences across its cell body and effectively migrate toward the chemoattractant source. Furthermore, it has been demonstrated in various studies that neutrophils are highly sensitive to changes in the surrounding chemoattractant environments, suggesting the role of a chemotactic memory for processing the complex spatiotemporal chemical guiding signals. Using a microfluidic device, in the present study we characterized neutrophil migration under spatially varying profiles of interleukine-8 <span class="hlt">gradients</span>, which consist of three spatially ordered regions of a shallow <span class="hlt">gradient</span>, a steep <span class="hlt">gradient</span> and a nearly saturated <span class="hlt">gradient</span>. This design allowed us to examine how neutrophils migrate under different chemoattractant <span class="hlt">gradient</span> profiles, and how the migratory response is affected when the cell moves from one <span class="hlt">gradient</span> profile to another in a single experiment. Our results show robust neutrophil chemotaxis in the shallow and steep <span class="hlt">gradient</span>, but not the saturated <span class="hlt">gradient</span>. Furthermore, neutrophils display a transition from chemotaxis to flowtaxis when they migrate across the steep <span class="hlt">gradient</span> interface, and the relative efficiency of this transition depends on the cell's chemotaxis history. Finally, some neutrophils were <span class="hlt">observed</span> to adjust their morphology to different <span class="hlt">gradient</span> profiles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDR12005A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDR12005A"><span><span class="hlt">Gradient</span> Augmented Level Set Method for Two Phase Flow Simulations with Phase Change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anumolu, C. R. Lakshman; Trujillo, Mario F.</p> <p>2016-11-01</p> <p>A sharp interface capturing approach is presented for two-phase flow simulations with phase change. The <span class="hlt">Gradient</span> Augmented Levelset method is coupled with the two-phase momentum and energy equations to advect the liquid-gas interface and predict heat transfer with phase change. The Ghost Fluid Method (GFM) is adopted for <span class="hlt">velocity</span> to discretize the advection and diffusion terms in the interfacial region. Furthermore, the GFM is employed to treat the discontinuity in the stress tensor, <span class="hlt">velocity</span>, and temperature <span class="hlt">gradient</span> yielding an accurate treatment in handling jump conditions. Thermal convection and diffusion terms are approximated by explicitly identifying the interface location, resulting in a sharp treatment for the energy solution. This sharp treatment is extended to estimate the interfacial mass transfer rate. At the computational cell, a d-cubic Hermite interpolating polynomial is employed to describe the interface location, which is locally fourth-order accurate. This extent of subgrid level description provides an accurate methodology for treating various interfacial processes with a high degree of sharpness. The ability to predict the interface and temperature evolutions accurately is illustrated by comparing numerical results with existing 1D to 3D analytical solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...841...13L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...841...13L"><span>Fluorescent H2 Emission Lines from the Reflection Nebula NGC 7023 <span class="hlt">Observed</span> with IGRINS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Le, Huynh Anh N.; Pak, Soojong; Kaplan, Kyle; Mace, Gregory; Lee, Sungho; Pavel, Michael; Jeong, Ueejeong; Oh, Heeyoung; Lee, Hye-In; Chun, Moo-Young; Yuk, In-Soo; Pyo, Tae-Soo; Hwang, Narae; Kim, Kang-Min; Park, Chan; Sok Oh, Jae; Yu, Young Sam; Park, Byeong-Gon; Minh, Young Chol; Jaffe, Daniel T.</p> <p>2017-05-01</p> <p>We have analyzed the temperature, <span class="hlt">velocity</span>, and density of H2 gas in NGC 7023 with a high-resolution near-infrared spectrum of the northwestern filament of the reflection nebula. By <span class="hlt">observing</span> NGC 7023 in the H and K bands at R ≃ 45,000 with the Immersion GRating INfrared Spectrograph, we detected 68 H2 emission lines within the 1″ × 15″ slit. The diagnostic ratio of 2-1 S(1)/1-0 S(1) is 0.41-0.56. In addition, the estimated ortho-to-para ratio (OPR) is 1.63-1.82, indicating that the H2 emission transitions in the <span class="hlt">observed</span> region arise mostly from gas excited by UV fluorescence. <span class="hlt">Gradients</span> in the temperature, <span class="hlt">velocity</span>, and OPR within the <span class="hlt">observed</span> area imply motion of the photodissociation region (PDR) relative to the molecular cloud. In addition, we derive the column density of H2 from the <span class="hlt">observed</span> emission lines and compare these results with PDR models in the literature covering a range of densities and incident UV field intensities. The notable difference between PDR model predictions and the <span class="hlt">observed</span> data, in high rotational J levels of ν = 1, is that the predicted formation temperature for newly formed H2 should be lower than that of the model predictions. To investigate the density distribution, we combine pixels in 1″ × 1″ areas and derive the density distribution at the 0.002 pc scale. The derived <span class="hlt">gradient</span> of density suggests that NGC 7023 has a clumpy structure, including a high clump density of ˜105 cm-3 with a size smaller than ˜5 × 10-3 pc embedded in lower-density regions of 103-104 cm-3.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.9284B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.9284B"><span>S-N secular ocean tide: explanation of <span class="hlt">observably</span> coastal <span class="hlt">velocities</span> of increase of a global mean sea level and mean sea levels in northern and southern hemispheres and prediction of erroneous altimetry <span class="hlt">velocities</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barkin, Yury</p> <p>2010-05-01</p> <p>The phenomenon of contrast secular changes of sea levels in the southern and northern hemispheres, predicted on the basis of geodynamic model about the forced relative oscillations and displacements of the Earth shells, has obtained a theoretical explanation. In northern hemisphere the mean sea level of ocean increases with <span class="hlt">velocity</span> about 2.45±0.32 mm/yr, and in a southern hemisphere the mean sea level increases with <span class="hlt">velocity</span> about 0.67±0.30 mm/yr. Theoretical values of <span class="hlt">velocity</span> of increase of global mean sea level of ocean has been estimated in 1.61±0.36 mm/yr. 1 Introduction. The secular drift of the centre of mass of the Earth in the direction of North Pole with <span class="hlt">velocity</span> about 12-20 mm/yr has been predicted by author in 1995 [1], [2], and now has confirmed with methods of space geodesy. For example the DORIS data in period 1999-2008 let us to estimate <span class="hlt">velocity</span> of polar drift in 5.24±0.29 mm/yr [3]. To explain this fundamental planetary phenomenon it is possible only, having admitted, that similar northern drift tests the centre of mass of the liquid core relatively to the centre of mass of viscous-elastic and thermodynamically changeable mantle with <span class="hlt">velocity</span> about 2-3 cm/yr in present [4]. The polar drift of the Earth core with huge superfluous mass results in slow increase of a gravity in northern hemisphere with a mean <span class="hlt">velocity</span> about 1.4 ?Gal and to its decrease approximately with the same mean <span class="hlt">velocity</span> in southern hemisphere [5]. This conclusion-prediction has obtained already a number of confirmations in precision gravimetric <span class="hlt">observations</span> fulfilled in last decade around the world [6]. Naturally, a drift of the core is accompanied by the global changes (deformations) of all layers of the mantle and the core, by inversion changes of their tension states when in one hemisphere the tension increases and opposite on the contrary - decreases. Also it is possible that thermodynamical mechanism actively works with inversion properties of molting and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663691-sdss-iv-mangarotation-velocity-lags-extraplanar-ionized-gas-from-manga-observations-edge-galaxies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663691-sdss-iv-mangarotation-velocity-lags-extraplanar-ionized-gas-from-manga-observations-edge-galaxies"><span>SDSS IV MaNGA—Rotation <span class="hlt">Velocity</span> Lags in the Extraplanar Ionized Gas from MaNGA <span class="hlt">Observations</span> of Edge-on Galaxies</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bizyaev, D.; Pan, K.; Brinkmann, J.</p> <p>2017-04-20</p> <p>We present a study of the kinematics of the extraplanar ionized gas around several dozen galaxies <span class="hlt">observed</span> by the Mapping of Nearby Galaxies at the Apache Point Observatory (MaNGA) survey. We considered a sample of 67 edge-on galaxies out of more than 1400 extragalactic targets <span class="hlt">observed</span> by MaNGA, in which we found 25 galaxies (or 37%) with regular lagging of the rotation curve at large distances from the galactic midplane. We model the <span class="hlt">observed</span> H α emission <span class="hlt">velocity</span> fields in the galaxies, taking projection effects and a simple model for the dust extinction into account. We show that the verticalmore » lag of the rotation curve is necessary in the modeling, and estimate the lag amplitude in the galaxies. We find no correlation between the lag and the star formation rate in the galaxies. At the same time, we report a correlation between the lag and the galactic stellar mass, central stellar <span class="hlt">velocity</span> dispersion, and axial ratio of the light distribution. These correlations suggest a possible higher ratio of infalling-to-local gas in early-type disk galaxies or a connection between lags and the possible presence of hot gaseous halos, which may be more prevalent in more massive galaxies. These results again demonstrate that <span class="hlt">observations</span> of extraplanar gas can serve as a potential probe for accretion of gas.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...839...87B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...839...87B"><span>SDSS IV MaNGA—Rotation <span class="hlt">Velocity</span> Lags in the Extraplanar Ionized Gas from MaNGA <span class="hlt">Observations</span> of Edge-on Galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bizyaev, D.; Walterbos, R. A. M.; Yoachim, P.; Riffel, R. A.; Fernández-Trincado, J. G.; Pan, K.; Diamond-Stanic, A. M.; Jones, A.; Thomas, D.; Cleary, J.; Brinkmann, J.</p> <p>2017-04-01</p> <p>We present a study of the kinematics of the extraplanar ionized gas around several dozen galaxies <span class="hlt">observed</span> by the Mapping of Nearby Galaxies at the Apache Point Observatory (MaNGA) survey. We considered a sample of 67 edge-on galaxies out of more than 1400 extragalactic targets <span class="hlt">observed</span> by MaNGA, in which we found 25 galaxies (or 37%) with regular lagging of the rotation curve at large distances from the galactic midplane. We model the <span class="hlt">observed</span> Hα emission <span class="hlt">velocity</span> fields in the galaxies, taking projection effects and a simple model for the dust extinction into account. We show that the vertical lag of the rotation curve is necessary in the modeling, and estimate the lag amplitude in the galaxies. We find no correlation between the lag and the star formation rate in the galaxies. At the same time, we report a correlation between the lag and the galactic stellar mass, central stellar <span class="hlt">velocity</span> dispersion, and axial ratio of the light distribution. These correlations suggest a possible higher ratio of infalling-to-local gas in early-type disk galaxies or a connection between lags and the possible presence of hot gaseous halos, which may be more prevalent in more massive galaxies. These results again demonstrate that <span class="hlt">observations</span> of extraplanar gas can serve as a potential probe for accretion of gas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SGeo...39..245N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SGeo...39..245N"><span>Spheroidal Integral Equations for Geodetic Inversion of Geopotential <span class="hlt">Gradients</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Novák, Pavel; Šprlák, Michal</p> <p>2018-03-01</p> <p>The static Earth's gravitational field has traditionally been described in geodesy and geophysics by the gravitational potential (geopotential for short), a scalar function of 3-D position. Although not directly <span class="hlt">observable</span>, geopotential functionals such as its first- and second-order <span class="hlt">gradients</span> are routinely measured by ground, airborne and/or satellite sensors. In geodesy, these <span class="hlt">observables</span> are often used for recovery of the static geopotential at some simple reference surface approximating the actual Earth's surface. A generalized mathematical model is represented by a surface integral equation which originates in solving Dirichlet's boundary-value problem of the potential theory defined for the harmonic geopotential, spheroidal boundary and globally distributed <span class="hlt">gradient</span> data. The mathematical model can be used for combining various geopotential <span class="hlt">gradients</span> without necessity of their re-sampling or prior continuation in space. The model extends the apparatus of integral equations which results from solving boundary-value problems of the potential theory to all geopotential <span class="hlt">gradients</span> <span class="hlt">observed</span> by current ground, airborne and satellite sensors. Differences between spherical and spheroidal formulations of integral kernel functions of Green's kind are investigated. Estimated differences reach relative values at the level of 3% which demonstrates the significance of spheroidal approximation for flattened bodies such as the Earth. The <span class="hlt">observation</span> model can be used for combined inversion of currently available geopotential <span class="hlt">gradients</span> while exploring their spectral and stochastic characteristics. The model would be even more relevant to gravitational field modelling of other bodies in space with more pronounced spheroidal geometry than that of the Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018A%26A...609A..12Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018A%26A...609A..12Z"><span>Spectrum radial <span class="hlt">velocity</span> analyser (SERVAL). High-precision radial <span class="hlt">velocities</span> and two alternative spectral indicators</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zechmeister, M.; Reiners, A.; Amado, P. J.; Azzaro, M.; Bauer, F. F.; Béjar, V. J. S.; Caballero, J. A.; Guenther, E. W.; Hagen, H.-J.; Jeffers, S. V.; Kaminski, A.; Kürster, M.; Launhardt, R.; Montes, D.; Morales, J. C.; Quirrenbach, A.; Reffert, S.; Ribas, I.; Seifert, W.; Tal-Or, L.; Wolthoff, V.</p> <p>2018-01-01</p> <p>Context. The CARMENES survey is a high-precision radial <span class="hlt">velocity</span> (RV) programme that aims to detect Earth-like planets orbiting low-mass stars. Aims: We develop least-squares fitting algorithms to derive the RVs and additional spectral diagnostics implemented in the SpEctrum Radial <span class="hlt">Velocity</span> AnaLyser (SERVAL), a publicly available python code. Methods: We measured the RVs using high signal-to-noise templates created by coadding all available spectra of each star. We define the chromatic index as the RV <span class="hlt">gradient</span> as a function of wavelength with the RVs measured in the echelle orders. Additionally, we computed the differential line width by correlating the fit residuals with the second derivative of the template to track variations in the stellar line width. Results: Using HARPS data, our SERVAL code achieves a RV precision at the level of 1 m/s. Applying the chromatic index to CARMENES data of the active star YZ CMi, we identify apparent RV variations induced by stellar activity. The differential line width is found to be an alternative indicator to the commonly used full width half maximum. Conclusions: We find that at the red optical wavelengths (700-900 nm) obtained by the visual channel of CARMENES, the chromatic index is an excellent tool to investigate stellar active regions and to identify and perhaps even correct for activity-induced RV variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012A%26A...537A.120Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012A%26A...537A.120Z"><span>Rotational <span class="hlt">velocities</span> of A-type stars. IV. Evolution of rotational <span class="hlt">velocities</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zorec, J.; Royer, F.</p> <p>2012-01-01</p> <p>Context. In previous works of this series, we have shown that late B- and early A-type stars have genuine bimodal distributions of rotational <span class="hlt">velocities</span> and that late A-type stars lack slow rotators. The distributions of the surface angular <span class="hlt">velocity</span> ratio Ω/Ωcrit (Ωcrit is the critical angular <span class="hlt">velocity</span>) have peculiar shapes according to spectral type groups, which can be caused by evolutionary properties. Aims: We aim to review the properties of these rotational <span class="hlt">velocity</span> distributions in some detail as a function of stellar mass and age. Methods: We have gathered vsini for a sample of 2014 B6- to F2-type stars. We have determined the masses and ages for these objects with stellar evolution models. The (Teff,log L/L⊙)-parameters were determined from the uvby-β photometry and the HIPPARCOS parallaxes. Results: The <span class="hlt">velocity</span> distributions show two regimes that depend on the stellar mass. Stars less massive than 2.5 M⊙ have a unimodal equatorial <span class="hlt">velocity</span> distribution and show a monotonical acceleration with age on the main sequence (MS). Stars more massive have a bimodal equatorial <span class="hlt">velocity</span> distribution. Contrarily to theoretical predictions, the equatorial <span class="hlt">velocities</span> of stars from about 1.7 M⊙ to 3.2 M⊙ undergo a strong acceleration in the first third of the MS evolutionary phase, while in the last third of the MS they evolve roughly as if there were no angular momentum redistribution in the external stellar layers. The studied stars might start in the ZAMS not necessarily as rigid rotators, but with a total angular momentum lower than the critical one of rigid rotators. The stars seem to evolve as differential rotators all the way of their MS life span and the variation of the <span class="hlt">observed</span> rotational <span class="hlt">velocities</span> proceeds with characteristic time scales δt ≈ 0.2 tMS, where tMS is the time spent by a star in the MS. Full Table 1 is only available at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.9303G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.9303G"><span>Effect of Anisotropic <span class="hlt">Velocity</span> Structure on Acoustic Emission Source Location during True-Triaxial Deformation Experiments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ghofrani Tabari, Mehdi; Goodfellow, Sebastian; Young, R. Paul</p> <p>2016-04-01</p> <p> transducer shots. AE of the rock during the whole experiment recorded by the surrounding transducers were investigated by location methods developed for anisotropic heterogeneous medium where, the M-shape fracture pattern was <span class="hlt">observed</span>. AE events occurred in the vicinity of the dilation pseudo-boundaries where, a relatively large <span class="hlt">velocity</span> <span class="hlt">gradient</span> was formed and along parallel fractures in the σ1/σ2 plane. This research is contributing to enhanced AE interpretation of fracture growth processes in the rock under laboratory true-triaxial stress conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20953374-toroidal-momentum-pinch-velocity-due-coriolis-drift-effect-small-scale-instabilities-toroidal-plasma','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20953374-toroidal-momentum-pinch-velocity-due-coriolis-drift-effect-small-scale-instabilities-toroidal-plasma"><span>Toroidal Momentum Pinch <span class="hlt">Velocity</span> due to the Coriolis Drift Effect on Small Scale Instabilities in a Toroidal Plasma</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Peeters, A. G.; Angioni, C.; Strintzi, D.</p> <p></p> <p>In this Letter, the influence of the ''Coriolis drift'' on small scale instabilities in toroidal plasmas is shown to generate a toroidal momentum pinch <span class="hlt">velocity</span>. Such a pinch results because the Coriolis drift generates a coupling between the density and temperature perturbations on the one hand and the perturbed parallel flow <span class="hlt">velocity</span> on the other. A simple fluid model is used to highlight the physics mechanism and gyro-kinetic calculations are performed to accurately assess the magnitude of the pinch. The derived pinch <span class="hlt">velocity</span> leads to a radial <span class="hlt">gradient</span> of the toroidal <span class="hlt">velocity</span> profile even in the absence of a torquemore » on the plasma and is predicted to generate a peaking of the toroidal <span class="hlt">velocity</span> profile similar to the peaking of the density profile. Finally, the pinch also affects the interpretation of current experiment000.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009EGUGA..11.6763B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009EGUGA..11.6763B"><span>Gravity <span class="hlt">gradient</span> preprocessing at the GOCE HPF</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bouman, J.; Rispens, S.; Gruber, T.; Schrama, E.; Visser, P.; Tscherning, C. C.; Veicherts, M.</p> <p>2009-04-01</p> <p>One of the products derived from the GOCE <span class="hlt">observations</span> are the gravity <span class="hlt">gradients</span>. These gravity <span class="hlt">gradients</span> are provided in the Gradiometer Reference Frame (GRF) and are calibrated in-flight using satellite shaking and star sensor data. In order to use these gravity <span class="hlt">gradients</span> for application in Earth sciences and gravity field analysis, additional pre-processing needs to be done, including corrections for temporal gravity field signals to isolate the static gravity field part, screening for outliers, calibration by comparison with existing external gravity field information and error assessment. The temporal gravity <span class="hlt">gradient</span> corrections consist of tidal and non-tidal corrections. These are all generally below the gravity <span class="hlt">gradient</span> error level, which is predicted to show a 1/f behaviour for low frequencies. In the outlier detection the 1/f error is compensated for by subtracting a local median from the data, while the data error is assessed using the median absolute deviation. The local median acts as a high-pass filter and it is robust as is the median absolute deviation. Three different methods have been implemented for the calibration of the gravity <span class="hlt">gradients</span>. All three methods use a high-pass filter to compensate for the 1/f gravity <span class="hlt">gradient</span> error. The baseline method uses state-of-the-art global gravity field models and the most accurate results are obtained if star sensor misalignments are estimated along with the calibration parameters. A second calibration method uses GOCE GPS data to estimate a low degree gravity field model as well as gravity <span class="hlt">gradient</span> scale factors. Both methods allow to estimate gravity <span class="hlt">gradient</span> scale factors down to the 10-3 level. The third calibration method uses high accurate terrestrial gravity data in selected regions to validate the gravity <span class="hlt">gradient</span> scale factors, focussing on the measurement band. Gravity <span class="hlt">gradient</span> scale factors may be estimated down to the 10-2 level with this method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PApGe.174.4459Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PApGe.174.4459Y"><span>Threshold <span class="hlt">Velocity</span> for Saltation Activity in the Taklimakan Desert</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Xinghua; He, Qing; Matimin, Ali; Yang, Fan; Huo, Wen; Liu, Xinchun; Zhao, Tianliang; Shen, Shuanghe</p> <p>2017-12-01</p> <p>The threshold <span class="hlt">velocity</span> is an indicator of a soil's susceptibility to saltation activity and is also an important parameter in dust emission models. In this study, the saltation activity, atmospheric conditions, and soil conditions were measured from 1 August 2008 to 31 July 2009 in the Taklimakan Desert, China. the threshold <span class="hlt">velocity</span> was estimated using the Gaussian time fraction equivalence method. At 2 m height, the 1-min averaged threshold <span class="hlt">velocity</span> varied between 3.5 and 10.9 m/s, with a mean of 5.9 m/s. Threshold <span class="hlt">velocities</span> varying between 4.5 and 7.5 m/s accounted for about 91.4% of all measurements. The average threshold <span class="hlt">velocity</span> displayed clear seasonal variations in the following sequence: winter (5.1 m/s) < autumn (5.8 m/s) < spring (6.1 m/s) < summer (6.5 m/s). A regression equation of threshold <span class="hlt">velocity</span> was established based on the relations between daily mean threshold <span class="hlt">velocity</span> and air temperature, specific humidity, and soil volumetric moisture content. High or moderate positive correlations were found between threshold <span class="hlt">velocity</span> and air temperature, specific humidity, and soil volumetric moisture content (air temperature r = 0.75; specific humidity r = 0.59; and soil volumetric moisture content r = 0.55; sample size = 251). In the study area, the <span class="hlt">observed</span> horizontal dust flux was 4198.0 kg/m during the whole period of <span class="hlt">observation</span>, while the horizontal dust flux calculated using the threshold <span class="hlt">velocity</span> from the regression equation was 4675.6 kg/m. The correlation coefficient between the calculated result and the <span class="hlt">observations</span> was 0.91. These results indicate that atmospheric and soil conditions should not be neglected in parameterization schemes for threshold <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.467..353M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.467..353M"><span>The imprints of bars on the vertical stellar population <span class="hlt">gradients</span> of galactic bulges</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molaeinezhad, A.; Falcón-Barroso, J.; Martínez-Valpuesta, I.; Khosroshahi, H. G.; Vazdekis, A.; La Barbera, F.; Peletier, R. F.; Balcells, M.</p> <p>2017-05-01</p> <p>This is the second paper of a series aimed to study the stellar kinematics and population properties of bulges in highly inclined barred galaxies. In this work, we carry out a detailed analysis of the stellar age, metallicity and [Mg/Fe] of 28 highly inclined (I > 65°) disc galaxies, from S0 to S(B)c, <span class="hlt">observed</span> with the SAURON integral-field spectrograph. The sample is divided into two clean samples of barred and unbarred galaxies, on the basis of the correlation between the stellar <span class="hlt">velocity</span> and h3 profiles, as well as the level of cylindrical rotation within the bulge region. We find that while the mean stellar age, metallicity and [Mg/Fe] in the bulges of barred and unbarred galaxies are not statistically distinct, the [Mg/Fe] <span class="hlt">gradients</span> along the minor axis (away from the disc) of barred galaxies are significantly different than those without bars. For barred galaxies, stars that are vertically further away from the mid-plane are in general more [Mg/Fe]-enhanced and thus the vertical <span class="hlt">gradients</span> in [Mg/Fe] for barred galaxies are mostly positive, while for unbarred bulges the [Mg/Fe] profiles are typically negative or flat. This result, together with the old populations <span class="hlt">observed</span> in the barred sample, indicates that bars are long-lasting structures, and therefore are not easily destroyed. The marked [Mg/Fe] differences with the bulges of unbarred galaxies indicate that different formation/evolution scenarios are required to explain their build-up, and emphasizes the role of bars in redistributing stellar material in the bulge-dominated regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSH41D2399T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSH41D2399T"><span>On the Anisotropy of the He+, C+, O+, and Ne+ Pickup Ion <span class="hlt">Velocity</span> Distribution Function: STEREO PLASTIC <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Taut, A.; Drews, C.; Berger, L.; Peleikis, T.; Wimmer-Schweingruber, R. F.</p> <p>2015-12-01</p> <p>PickUp Ions (PUIs) are typically characterized by (1) their almost exclusively single charge state, (2) a highly non-thermal and anisotropic <span class="hlt">Velocity</span> Distribution Function (VDF) [Drews et al., 2015], and (3) an extended source population of neutral atoms somewhere between the <span class="hlt">observer</span> and the Sun. The origin of pickup ions ranges from sources only several solar radii away from the Sun, the so-called inner-source of pickup ions, up to a distance of several hundreds of astronomical units, the local interstellar medium. Their continuous production inside the heliosphere and complex interactions with the magnetized solar wind plasma leads to the development of non-thermal, anisotropic features of both the solar wind and pickup ion <span class="hlt">velocity</span> distribution functions. In this study, we present <span class="hlt">observations</span> of the VDF of He+, C+, N+, O+ and Ne+ pickup ions with PLASTIC on STEREO A. We have found a PUI flux increase during perpendicular configurations of the local magnetic field that is generally linked to the existence of a so-called torus-distribution [Drews et al., 2015] which is attributed to the production of PUIs close to the <span class="hlt">observer</span>. A comparison of the PUI VDF between radial and perpendicular configurations of the local magnetic field vector is used to quantify the anisotropy of the PUI VDF and thereby enables us to estimate the mean free path for pitch-angle scattering of He, C, N, O and Ne pickup ions without the necessity of an over-simplified heliospheric model to describe the PUI phase space transport. Our results show a clear signature of a C+ torus signature at 1 AU as well as significant differences between the anisotropies of the He+ and O+ VDF. We will discuss our results in the light of recent studies about the nature of the inner-source of PUIs [Berger et al., 2015] and <span class="hlt">observations</span> of the 2D VDF of He+[Drews et al., 2015]. Figure Caption: <span class="hlt">Velocity</span> space diagrams of a pickup ion torus distribution as a (vx-vy)-projection (top left panel) and in the vz = 0</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17399733','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17399733"><span>Application of a chromatography model with linear <span class="hlt">gradient</span> elution experimental data to the rapid scale-up in ion-exchange process chromatography of proteins.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ishihara, Takashi; Kadoya, Toshihiko; Yamamoto, Shuichi</p> <p>2007-08-24</p> <p>We applied the model described in our previous paper to the rapid scale-up in the ion exchange chromatography of proteins, in which linear flow <span class="hlt">velocity</span>, column length and <span class="hlt">gradient</span> slope were changed. We carried out linear <span class="hlt">gradient</span> elution experiments, and obtained data for the peak salt concentration and peak width. From these data, the plate height (HETP) was calculated as a function of the mobile phase <span class="hlt">velocity</span> and iso-resolution curve (the separation time and elution volume relationship for the same resolution) was calculated. The scale-up chromatography conditions were determined by the iso-resolution curve. The scale-up of the linear <span class="hlt">gradient</span> elution from 5 to 100mL and 2.5L column sizes was performed both by the separation of beta-lactoglobulin A and beta-lactoglobulin B with anion-exchange chromatography and by the purification of a recombinant protein with cation-exchange chromatography. Resolution, recovery and purity were examined in order to verify the proposed method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.8374G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.8374G"><span>First high resolution P wave <span class="hlt">velocity</span> structure beneath Tenerife Island, (Canary Islands, Spain)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Garcia-Yeguas, Araceli; Ivan, Koulakov; Ibañez Jesus, M.; Valenti, Sallarès.</p> <p>2010-05-01</p> <p>3D <span class="hlt">velocity</span> structure distribution has been imaged for first time using high resolution traveltime seismic tomography of the active volcano of Tenerife Island (Canary Islands, Spain). It is located in the Atlantic Ocean. In this island is situated the Teide stratovolcano (3718 m high) that is part of the Cañadas-Teide-Pico Viejo volcanic complex. Las Cañadas is a caldera system more than 20 kilometers wide where at least four distinct caldera processes have been identified. Evidence for many explosive eruptions in the volcanic complex has been found; the last noticeable explosive eruption (sub-plinean) occurred at Montaña Blanca around 2000 years ago. During the last 300 years, six effusive eruptions have been reported, the last of which took place at Chinyero Volcano on 18 November 1909. In January 2007, a seismic active experiment was carried out as part of the TOM-TEIDEVS project. About 6850 air gun shots were fired on the sea and recorded on a dense local seismic land network consisting of 150 independent (three component) seismic stations. The good quality of the recorded data allowed identifying P-wave arrivals up to offsets of 30-40 km obtaining more than 63000 traveltimes used in the tomographic inversion. The images have been obtained using ATOM-3D code (Koulakov, 2009). This code uses ray bending algorithms in the ray tracing for the forward modelling and in the inversion step it uses <span class="hlt">gradient</span> methods. The <span class="hlt">velocity</span> models show a very heterogeneous upper crust that is usual in similar volcanic environment. The tomographic images points out the no-existence of a magmatic chamber near to the surface and below Pico Teide. The ancient Las Cañadas caldera borders are clearly imaged featuring relatively high seismic <span class="hlt">velocity</span>. Moreover, we have found a big low <span class="hlt">velocity</span> anomaly in the northwest dorsal of the island. The last eruption took place in 1909 in this area. Furthermore, in the southeast another low <span class="hlt">velocity</span> anomaly has been imaged. Several resolution</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25502599','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25502599"><span>Characterization of the startup transient electrokinetic flow in rectangular channels of arbitrary dimensions, zeta potential distribution, and time-varying pressure <span class="hlt">gradient</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Miller, Andrew; Villegas, Arturo; Diez, F Javier</p> <p>2015-03-01</p> <p>The solution to the startup transient EOF in an arbitrary rectangular microchannel is derived analytically and validated experimentally. This full 2D transient solution describes the evolution of the flow through five distinct periods until reaching a final steady state. The derived analytical <span class="hlt">velocity</span> solution is validated experimentally for different channel sizes and aspect ratios under time-varying pressure <span class="hlt">gradients</span>. The experiments used a time resolved micro particle image velocimetry technique to calculate the startup transient <span class="hlt">velocity</span> profiles. The measurements captured the effect of time-varying pressure <span class="hlt">gradient</span> fields derived in the analytical solutions. This is tested by using small reservoirs at both ends of the channel which allowed a time-varying pressure <span class="hlt">gradient</span> to develop with a time scale on the order of the transient EOF. Results showed that under these common conditions, the effect of the pressure build up in the reservoirs on the temporal development of the transient startup EOF in the channels cannot be neglected. The measurements also captured the analytical predictions for channel walls made of different materials (i.e., zeta potentials). This was tested in channels that had three PDMS and one quartz wall, resulting in a flow with an asymmetric <span class="hlt">velocity</span> profile due to variations in the zeta potential between the walls. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930062338&hterms=heinz&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dheinz','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930062338&hterms=heinz&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dheinz"><span>A Doppler dimming determination of coronal outflow <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Strachan, Leonard; Kohl, John L.; Weiser, Heinz; Withbroe, George L.; Munro, Richard H.</p> <p>1993-01-01</p> <p>Outflow <span class="hlt">velocities</span> in a polar coronal hole are derived from <span class="hlt">observations</span> made during a 1982 sounding rocket flight. The <span class="hlt">velocity</span> results are derived from a Doppler dimming analysis of resonantly scattered H I Ly-alpha. This analysis indicates radial outflow <span class="hlt">velocities</span> of 217 km/s at 2 solar radii from sun-center with an uncertainty range of 153 to 251 km/s at a confidence level of 67 percent. These results are best characterized as strong evidence for supersonic outflow within 2 solar radii of sun-center in a polar coronal hole. Several means for obtaining improved accuracy in future <span class="hlt">observations</span> are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192861','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192861"><span>Field and laboratory determination of water-surface elevation and <span class="hlt">velocity</span> using noncontact measurements</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nelson, Jonathan M.; Kinzel, Paul J.; Schmeeckle, Mark Walter; McDonald, Richard R.; Minear, Justin T.</p> <p>2016-01-01</p> <p>Noncontact methods for measuring water-surface elevation and <span class="hlt">velocity</span> in laboratory flumes and rivers are presented with examples. Water-surface elevations are measured using an array of acoustic transducers in the laboratory and using laser scanning in field situations. Water-surface <span class="hlt">velocities</span> are based on using particle image velocimetry or other machine vision techniques on infrared video of the water surface. Using spatial and temporal averaging, results from these methods provide information that can be used to develop estimates of discharge for flows over known bathymetry. Making such estimates requires relating water-surface <span class="hlt">velocities</span> to vertically averaged <span class="hlt">velocities</span>; the methods here use standard relations. To examine where these relations break down, laboratory data for flows over simple bumps of three amplitudes are evaluated. As anticipated, discharges determined from surface information can have large errors where nonhydrostatic effects are large. In addition to investigating and characterizing this potential error in estimating discharge, a simple method for correction of the issue is presented. With a simple correction based on bed <span class="hlt">gradient</span> along the flow direction, remotely sensed estimates of discharge appear to be viable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMNS23A1907T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMNS23A1907T"><span>Use of Ground Penetrating Radar to Study <span class="hlt">Gradient</span> Media</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Titov, A.</p> <p>2016-12-01</p> <p>Nowadays Ground Penetrating Radar (GPR) is often used to solve different problems of applied geophysics including the hydrological ones. This work was motivated by detection of weak reflections in the body of water <span class="hlt">observed</span> during the surveys on the freshwater lakes using GPR. The same reflections were first analyzed by John Bradford in 2007. These reflections can arise from the thermal <span class="hlt">gradient</span> layer or thermocline due to different dielectric permittivity of cold and warm water. We employed physical and mathematical modeling to identify the properties of such thermoclines. We have constructed a special GPR stand to study the <span class="hlt">gradient</span> media in our laboratory. The stand consists of a water-filled plastic tank and plastic tubes, which gather the cold water under the warm water. Our stand allows for changing parameters of the <span class="hlt">gradient</span> layer, such as limits of dielectric permittivity and the thickness of the <span class="hlt">gradient</span> layer. GPR antenna was placed slightly under the water surface to remove the parasitic reflections. To visualize the thermal distribution, an infrared camera and thermal sensors were used. Analysis of the GPR traces after physical modeling, performed in the MATLAB environment, allows us to locate the weak reflection from the <span class="hlt">gradient</span> layer. We <span class="hlt">observed</span> that (i) the change of the <span class="hlt">gradient</span> boundary values alters the amplitude of the signal, (ii) the arrival time of the impulse reflected from the <span class="hlt">gradient</span> layer corresponds to the arrival time of the impulse reflected from the top boundary of this layer, and (iii) the shape of the signal reflected from the <span class="hlt">gradient</span> layer coincides with the shape of the signal reflected from the non-<span class="hlt">gradient</span> boundary between two bodies. The quantitative properties of thermocline can be determined using amplitude analysis of GPR signals. Finally, the developed methods were successfully applied to real field data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T33D0765H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T33D0765H"><span>Detailed seismic <span class="hlt">velocity</span> structure of the ultra-slow spread crust at the Mid-Cayman Spreading Center from travel-time tomography and synthetic seismograms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harding, J.; Van Avendonk, H. J.; Hayman, N. W.; Grevemeyer, I.; Peirce, C.</p> <p>2017-12-01</p> <p>The Mid-Cayman Spreading Center (MCSC), an ultraslow-spreading center in the Caribbean Sea, has formed highly variable oceanic crust. Seafloor dredges have recovered extrusive basalts in the axial deeps as well as gabbro on bathymetric highs and exhumed mantle peridotite along the only 110 km MCSC. Wide-angle refraction data were collected with active-source ocean bottom seismometers in April, 2015, along lines parallel and across the MCSC. Travel-time tomography produces relatively smooth 2-D tomographic models of compressional wave <span class="hlt">velocity</span>. These <span class="hlt">velocity</span> models reveal large along- and across-axis variations in seismic <span class="hlt">velocity</span>, indicating possible changes in crustal thickness, composition, faulting, and magmatism. It is difficult, however, to differentiate between competing interpretations of seismic <span class="hlt">velocity</span> using these tomographic models alone. For example, in some areas the seismic <span class="hlt">velocities</span> may be explained by either thin igneous crust or exhumed, serpentinized mantle. Distinguishing between these two interpretations is important as we explore the relationships between magmatism, faulting, and hydrothermal venting at ultraslow-spreading centers. We therefore improved our constraints on the shallow seismic <span class="hlt">velocity</span> structure of the MCSC by modeling the amplitude of seismic refractions in the wide-angle data set. Synthetic seismograms were calculated with a finite-difference method for a range of models with different vertical <span class="hlt">velocity</span> <span class="hlt">gradients</span>. Small-scale features in the <span class="hlt">velocity</span> models, such as steep <span class="hlt">velocity</span> <span class="hlt">gradients</span> and Moho boundaries, were explored systematically to best fit the real data. With this approach, we have improved our understanding of the compressional <span class="hlt">velocity</span> structure of the MCSC along with the geological interpretations that are consistent with three seismic refraction profiles. Line P01 shows a variation in the thinness of lower seismic <span class="hlt">velocities</span> along the axis, indicating two segment centers, while across-axis lines P02 and P03</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29551240','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29551240"><span><span class="hlt">Gradient</span> elution behavior of proteins in hydrophobic interaction chromatography with U-shaped retention factor curves.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Creasy, Arch; Lomino, Joseph; Barker, Gregory; Khetan, Anurag; Carta, Giorgio</p> <p>2018-04-27</p> <p>Protein retention in hydrophobic interaction chromatography is described by the solvophobic theory as a function of the kosmostropic salt concentration. In general, an increase in salt concentration drives protein partitioning to the hydrophobic surface while a decrease reduces it. In some cases, however, protein retention also increases at low salt concentrations resulting in a U-shaped retention factor curve. During <span class="hlt">gradient</span> elution the salt concentration is gradually decreased from a high value thereby reducing the retention factor and increasing the protein chromatographic <span class="hlt">velocity</span>. For these conditions, a steep <span class="hlt">gradient</span> can overtake the protein in the column, causing it to rebind. Two dynamic models, one based on the local equilibrium theory and the other based on the linear driving force approximation, are presented. We show that the normalized <span class="hlt">gradient</span> slope determines whether the protein elutes in the <span class="hlt">gradient</span>, partially elutes, or is trapped in the column. Experimental results are presented for two different monoclonal antibodies and for lysozyme on Capto Phenyl (High Sub) resin. One of the mAbs and lysozyme exhibit U-shaped retention factor curves and for each, we determine the critical <span class="hlt">gradient</span> slope beyond which 100% recovery is no longer possible. Elution with a reverse <span class="hlt">gradient</span> is also demonstrated at low salt concentrations for these proteins. Understanding this behavior has implications in the design of <span class="hlt">gradient</span> elution since the <span class="hlt">gradient</span> slope impacts protein recovery. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19780029799&hterms=copernicus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcopernicus','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780029799&hterms=copernicus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dcopernicus"><span>Long-term changes in ultraviolet P Cygni profiles <span class="hlt">observed</span> with Copernicus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Snow, T. P., Jr.</p> <p>1977-01-01</p> <p>The incidence and nature of variability occurring on time scales of years in the ultraviolet P Cygni profiles of 15 O and B stars are investigated using spectrophotometric data obtained with the Copernicus satellite. It is found that some change in at least a few details of the P Cygni profiles is evident in almost every case, that the changes in a few stars appear to represent substantial variations in the column densities of the particular ions <span class="hlt">observed</span>, and that the changes in other stars are minor in nature and do not result from significant alterations in the quantity of material in the stellar winds. Most of the narrow absorption features are shown to be invariant in <span class="hlt">velocity</span>, although their strengths have apparently changed in certain cases. The nature of the changes <span class="hlt">observed</span> in each of the program stars is briefly described, the time scale for variability in the stellar winds is considered, and two stars (Zeta Pup and Delta Ori A) are identified for which some alteration in the total amount of material in the stellar wind has taken place. It is suggested that the narrow absorption features probably represent temperature <span class="hlt">gradients</span> or plateaus in the stellar-wind <span class="hlt">velocity</span> fields or may be caused by flat regions in the height dependence of the wind <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.S41C2208C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.S41C2208C"><span>Seismic <span class="hlt">velocity</span> variations at the EGS geothermal reservoir of Soultz-Sous-Forêts (France): Some <span class="hlt">observations</span> for understanding stress regime changes during hydraulic stimulations using 4D tomography</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Calo, M.; Dorbath, C.; Cornet, F.; Cuenot, N.</p> <p>2011-12-01</p> <p>During the last decade three deep wells (GPK2, GPK3, and GPK4) were drilled to a depth of about 5000 m at the Enhanced Geothermal System (EGS) site of Soultz-sous-Forêts (Alsace, France). All the wells were stimulated through high-pressure hydraulic injections. Several thousands of micro-earthquakes with Duration Magnitude ranging from -0.9 to 2.9 were produced. The induced earthquakes were located by downhole and surface seismic stations. The wells behaved differently during and after the stimulations, as shown by several authors. We present here a comparison between new 4D seismic tomographies performed for the above mentioned stimulation tests. The <span class="hlt">velocity</span> models have been obtained using the Double-Difference tomographic method (Zhang and Thurber 2003) and have been further improved with the post-processing WAM technique (Calo' et al., 2009, 2011). For each stimulation test, the subsetting of the data was performed by taking into account injection parameters (the injected flow rate and the wellhead pressure). In this work we discuss some important steps <span class="hlt">observed</span> during and after the injections. A first <span class="hlt">observation</span> is that low <span class="hlt">velocity</span> anomalies were centered around the wells when stimulations started and then disappeared just after strong changes in the injected flow rate. We interpret these changes in seismic properties as transient changes in the stress regime during the stimulations. Furthermore, as shown by the seismic <span class="hlt">velocity</span> models, pre-existing fracture network played a fundamental role on the intensity and distribution of the <span class="hlt">observed</span> <span class="hlt">velocity</span> anomalies. Indeed we <span class="hlt">observe</span> that low <span class="hlt">velocity</span> anomalies are much less evident and moved away from the well when documented large pre-exiting fractures cross the openhole part of the well. In particular, we <span class="hlt">observed</span> this pattern for the models calculated with the data of the GPK3 stimulation. Thanks to the improvement and the reliability of these new <span class="hlt">velocity</span> models, new discussions about the mechanical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.1625L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.1625L"><span>Six-hourly time series of horizontal troposphere <span class="hlt">gradients</span> in VLBI analyis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Landskron, Daniel; Hofmeister, Armin; Mayer, David; Böhm, Johannes</p> <p>2016-04-01</p> <p>Consideration of horizontal <span class="hlt">gradients</span> is indispensable for high-precision VLBI and GNSS analysis. As a rule of thumb, all <span class="hlt">observations</span> below 15 degrees elevation need to be corrected for the influence of azimuthal asymmetry on the delay times, which is mainly a product of the non-spherical shape of the atmosphere and ever-changing weather conditions. Based on the well-known <span class="hlt">gradient</span> estimation model by Chen and Herring (1997), we developed an augmented <span class="hlt">gradient</span> model with additional parameters which are determined from ray-traced delays for the complete history of VLBI <span class="hlt">observations</span>. As input to the ray-tracer, we used operational and re-analysis data from the European Centre for Medium-Range Weather Forecasts. Finally, we applied those a priori <span class="hlt">gradient</span> parameters to VLBI analysis along with other empirical <span class="hlt">gradient</span> models and assessed their impact on baseline length repeatabilities as well as on celestial and terrestrial reference frames.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ISPAr42W4..517R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ISPAr42W4..517R"><span>Co-Seismic Gravity <span class="hlt">Gradient</span> Changes of the 2006-2007 Great Earthquakes in the Central Kuril Islands from GRACE <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rahimi, A.; Shahrisvand, M.</p> <p>2017-09-01</p> <p>GRACE satellites (the Gravity Recovery And climate Experiment) are very useful sensors to extract gravity anomalies after earthquakes. In this study, we reveal co-seismic signals of the two combined earthquakes, the 2006 Mw8.3 thrust and 2007 Mw8.1 normal fault earthquakes of the central Kuril Islands from GRACE <span class="hlt">observations</span>. We compute monthly full gravitational <span class="hlt">gradient</span> tensor in the local north-east-down frame for Kuril Islands earthquakes without spatial averaging and de-striping filters. Some of gravitational <span class="hlt">gradient</span> components (e.g. ΔVxx, ΔVxz) enhance high frequency components of the earth gravity field and reveal more details in spatial and temporal domain. Therefore, co-seismic activity can be better illustrated. For the first time, we show that the positive-negative-positive co-seismic ΔVxx due to the Kuril Islands earthquakes ranges from - 0.13 to + 0.11 milli Eötvös, and ΔVxz shows a positive-negative-positive pattern ranges from - 0.16 to + 0.13 milli Eötvös, agree well with seismic model predictions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013RaSc...48..659B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013RaSc...48..659B"><span>Methodology of automated ionosphere front <span class="hlt">velocity</span> estimation for ground-based augmentation of GNSS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bang, Eugene; Lee, Jiyun</p> <p>2013-11-01</p> <p>ionospheric anomalies occurring during severe ionospheric storms can pose integrity threats to Global Navigation Satellite System (GNSS) Ground-Based Augmentation Systems (GBAS). Ionospheric anomaly threat models for each region of operation need to be developed to analyze the potential impact of these anomalies on GBAS users and develop mitigation strategies. Along with the magnitude of ionospheric <span class="hlt">gradients</span>, the speed of the ionosphere "fronts" in which these <span class="hlt">gradients</span> are embedded is an important parameter for simulation-based GBAS integrity analysis. This paper presents a methodology for automated ionosphere front <span class="hlt">velocity</span> estimation which will be used to analyze a vast amount of ionospheric data, build ionospheric anomaly threat models for different regions, and monitor ionospheric anomalies continuously going forward. This procedure automatically selects stations that show a similar trend of ionospheric delays, computes the orientation of detected fronts using a three-station-based trigonometric method, and estimates speeds for the front using a two-station-based method. It also includes fine-tuning methods to improve the estimation to be robust against faulty measurements and modeling errors. It demonstrates the performance of the algorithm by comparing the results of automated speed estimation to those manually computed previously. All speed estimates from the automated algorithm fall within error bars of ± 30% of the manually computed speeds. In addition, this algorithm is used to populate the current threat space with newly generated threat points. A larger number of <span class="hlt">velocity</span> estimates helps us to better understand the behavior of ionospheric <span class="hlt">gradients</span> under geomagnetic storm conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003PEPI..140..219G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003PEPI..140..219G"><span>D″ shear <span class="hlt">velocity</span> heterogeneity, anisotropy and discontinuity structure beneath the Caribbean and Central America</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Garnero, Edward J.; Lay, Thorne</p> <p>2003-11-01</p> <p> thickness or <span class="hlt">velocity</span> <span class="hlt">gradients</span> of the high-<span class="hlt">velocity</span> layer. While small-scale geographic patterns of heterogeneity, anisotropy, and discontinuity are present, the details appear complex, and require higher resolution array analyses to fully characterize the structure. Explanations for the high-shear wave speeds, anisotropy, and reflector associated with D″ beneath the Caribbean and Central America must be applicable over a lateral scale of roughly 1500 km 2, the dimension over which we <span class="hlt">observe</span> coherent wavefield behavior in the region. A slab graveyard appears viable in this regard.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhDT.......183S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhDT.......183S"><span><span class="hlt">Observing</span> Semi-Arid Ecoclimates across Mountain <span class="hlt">Gradients</span> in the Great Basin, USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Strachan, Scotty</p> <p></p> <p><span class="hlt">Observation</span> of climate and ecohydrological variables in mountain systems is a necessary (if challenging) endeavor for modern society. Water resources are often intimately tied to mountains, and high elevation environments are frequently home to unique landscapes and biota with limited geographical distributions. This is especially true in the temperate and semi-arid mountains of the western United States, and specifically the Great Basin. Stark contrasts in annual water balance and ecological populations are visible across steep elevational <span class="hlt">gradients</span> in the region; and yet the bulk of our historical knowledge of climate and related processes comes from lowland <span class="hlt">observations</span>. Interpolative models that strive to estimate conditions in mountains using existing datasets are often found to be inaccurate, making future projections of mountain climate and ecosystem response suspect. This study details the results of high-resolution topographically-diverse ecohydrological monitoring, and describes the character and seasonality of basic climatic variables such as temperature and precipitation as well as their impact on soil moisture and vegetation during the 2012-2015 drought sequence. Relationships of topography (elevation/aspect) to daily and seasonal temperatures are shown. Tests of the PRISM temperature model are performed at the large watershed scale, revealing magnitudes, modes, and potential sources of bias that could dramatically affect derivative scientific conclusions. A new method of precipitation phase partitioning to detect and quantify frozen precipitation on a sub-daily basis is described. Character of precipitation from sub-daily to annual scales is quantified across all major Great Basin vegetation/elevation zones, and the relationship of elevation to precipitation phase, intensity, and amount is explored. Water-stress responses of Great Basin conifers including Pinus flexilis, Pinus longaeva, and Pinus ponderosa are directly <span class="hlt">observed</span>, showing potential</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663600-fluorescent-sub-emission-lines-from-reflection-nebula-ngc-observed-igrins','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663600-fluorescent-sub-emission-lines-from-reflection-nebula-ngc-observed-igrins"><span>Fluorescent H{sub 2} Emission Lines from the Reflection Nebula NGC 7023 <span class="hlt">Observed</span> with IGRINS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Le, Huynh Anh N.; Pak, Soojong; Lee, Hye-In</p> <p></p> <p>We have analyzed the temperature, <span class="hlt">velocity</span>, and density of H{sub 2} gas in NGC 7023 with a high-resolution near-infrared spectrum of the northwestern filament of the reflection nebula. By <span class="hlt">observing</span> NGC 7023 in the H and K bands at R ≃ 45,000 with the Immersion GRating INfrared Spectrograph, we detected 68 H{sub 2} emission lines within the 1″ × 15″ slit. The diagnostic ratio of 2-1 S(1)/1-0 S(1) is 0.41−0.56. In addition, the estimated ortho-to-para ratio (OPR) is 1.63−1.82, indicating that the H{sub 2} emission transitions in the <span class="hlt">observed</span> region arise mostly from gas excited by UV fluorescence. <span class="hlt">Gradients</span> inmore » the temperature, <span class="hlt">velocity</span>, and OPR within the <span class="hlt">observed</span> area imply motion of the photodissociation region (PDR) relative to the molecular cloud. In addition, we derive the column density of H{sub 2} from the <span class="hlt">observed</span> emission lines and compare these results with PDR models in the literature covering a range of densities and incident UV field intensities. The notable difference between PDR model predictions and the <span class="hlt">observed</span> data, in high rotational J levels of ν = 1, is that the predicted formation temperature for newly formed H{sub 2} should be lower than that of the model predictions. To investigate the density distribution, we combine pixels in 1″ × 1″ areas and derive the density distribution at the 0.002 pc scale. The derived <span class="hlt">gradient</span> of density suggests that NGC 7023 has a clumpy structure, including a high clump density of ∼10{sup 5} cm{sup −3} with a size smaller than ∼5 × 10{sup −3} pc embedded in lower-density regions of 10{sup 3}–10{sup 4} cm{sup −3}.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22666224-observational-detection-drift-velocity-between-ionized-neutral-species-solar-prominences','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22666224-observational-detection-drift-velocity-between-ionized-neutral-species-solar-prominences"><span><span class="hlt">OBSERVATIONAL</span> DETECTION OF DRIFT <span class="hlt">VELOCITY</span> BETWEEN IONIZED AND NEUTRAL SPECIES IN SOLAR PROMINENCES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Khomenko, Elena; Collados, Manuel; Díaz, Antonio J., E-mail: khomenko@iac.es, E-mail: mcv@iac.es, E-mail: aj.diaz@uib.es</p> <p>2016-06-01</p> <p>We report the detection of differences in the ion and neutral <span class="hlt">velocities</span> in prominences using high-resolution spectral data obtained in 2012 September at the German Vacuum Tower Telescope (Observatorio del Teide, Tenerife). A time series of scans of a small portion of a solar prominence was obtained simultaneously with high cadence using the lines of two elements with different ionization states, namely, Ca ii 8542 Å and He i 10830 Å. The displacements, widths, and amplitudes of both lines were carefully compared to extract dynamical information about the plasma. Many dynamical features are detected, such as counterstreaming flows, jets, andmore » propagating waves. In all of the cases, we find a very strong correlation between the parameters extracted from the lines of both elements, confirming that both lines trace the same plasma. Nevertheless, we also find short-lived transients where this correlation is lost. These transients are associated with ion-neutral drift <span class="hlt">velocities</span> of the order of several hundred m s{sup −1}. The patches of non-zero drift <span class="hlt">velocity</span> show coherence in time–distance diagrams.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...597A..73N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...597A..73N"><span>HARPS-N high spectral resolution <span class="hlt">observations</span> of Cepheids I. The Baade-Wesselink projection factor of δ Cep revisited</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nardetto, N.; Poretti, E.; Rainer, M.; Fokin, A.; Mathias, P.; Anderson, R. I.; Gallenne, A.; Gieren, W.; Graczyk, D.; Kervella, P.; Mérand, A.; Mourard, D.; Neilson, H.; Pietrzynski, G.; Pilecki, B.; Storm, J.</p> <p>2017-01-01</p> <p>Context. The projection factor p is the key quantity used in the Baade-Wesselink (BW) method for distance determination; it converts radial <span class="hlt">velocities</span> into pulsation <span class="hlt">velocities</span>. Several methods are used to determine p, such as geometrical and hydrodynamical models or the inverse BW approach when the distance is known. Aims: We analyze new HARPS-N spectra of δ Cep to measure its cycle-averaged atmospheric <span class="hlt">velocity</span> <span class="hlt">gradient</span> in order to better constrain the projection factor. Methods: We first apply the inverse BW method to derive p directly from <span class="hlt">observations</span>. The projection factor can be divided into three subconcepts: (1) a geometrical effect (p0); (2) the <span class="hlt">velocity</span> <span class="hlt">gradient</span> within the atmosphere (fgrad); and (3) the relative motion of the optical pulsating photosphere with respect to the corresponding mass elements (fo-g). We then measure the fgrad value of δ Cep for the first time. Results: When the HARPS-N mean cross-correlated line-profiles are fitted with a Gaussian profile, the projection factor is pcc-g = 1.239 ± 0.034(stat.) ± 0.023(syst.). When we consider the different amplitudes of the radial <span class="hlt">velocity</span> curves that are associated with 17 selected spectral lines, we measure projection factors ranging from 1.273 to 1.329. We find a relation between fgrad and the line depth measured when the Cepheid is at minimum radius. This relation is consistent with that obtained from our best hydrodynamical model of δ Cep and with our projection factor decomposition. Using the <span class="hlt">observational</span> values of p and fgrad found for the 17 spectral lines, we derive a semi-theoretical value of fo-g. We alternatively obtain fo-g = 0.975 ± 0.002 or 1.006 ± 0.002 assuming models using radiative transfer in plane-parallel or spherically symmetric geometries, respectively. Conclusions: The new HARPS-N <span class="hlt">observations</span> of δ Cep are consistent with our decomposition of the projection factor. The next step will be to measure p0 directly from the next generation of visible interferometers</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25978093','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25978093"><span>Combining Step <span class="hlt">Gradients</span> and Linear <span class="hlt">Gradients</span> in Density.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kumar, Ashok A; Walz, Jenna A; Gonidec, Mathieu; Mace, Charles R; Whitesides, George M</p> <p>2015-06-16</p> <p>Combining aqueous multiphase systems (AMPS) and magnetic levitation (MagLev) provides a method to produce hybrid <span class="hlt">gradients</span> in apparent density. AMPS—solutions of different polymers, salts, or surfactants that spontaneously separate into immiscible but predominantly aqueous phases—offer thermodynamically stable steps in density that can be tuned by the concentration of solutes. MagLev—the levitation of diamagnetic objects in a paramagnetic fluid within a magnetic field gradient—can be arranged to provide a near-linear <span class="hlt">gradient</span> in effective density where the height of a levitating object above the surface of the magnet corresponds to its density; the strength of the <span class="hlt">gradient</span> in effective density can be tuned by the choice of paramagnetic salt and its concentrations and by the strength and <span class="hlt">gradient</span> in the magnetic field. Including paramagnetic salts (e.g., MnSO4 or MnCl2) in AMPS, and placing them in a magnetic field <span class="hlt">gradient</span>, enables their use as media for MagLev. The potential to create large steps in density with AMPS allows separations of objects across a range of densities. The <span class="hlt">gradients</span> produced by MagLev provide resolution over a continuous range of densities. By combining these approaches, mixtures of objects with large differences in density can be separated and analyzed simultaneously. Using MagLev to add an effective <span class="hlt">gradient</span> in density also enables tuning the range of densities captured at an interface of an AMPS by simply changing the position of the container in the magnetic field. Further, by creating AMPS in which phases have different concentrations of paramagnetic ions, the phases can provide different resolutions in density. These results suggest that combining steps in density with <span class="hlt">gradients</span> in density can enable new classes of separations based on density.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29034403','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29034403"><span>Customised spatiotemporal temperature <span class="hlt">gradients</span> created by a liquid metal enabled vortex generator.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhu, Jiu Yang; Thurgood, Peter; Nguyen, Ngan; Ghorbani, Kamran; Khoshmanesh, Khashayar</p> <p>2017-11-07</p> <p>Generating customised temperature <span class="hlt">gradients</span> in miniaturised flow-free liquid chambers is challenging due to the dominance of diffusion. Inducing internal flows in the form of vortices is an effective strategy for overcoming the limitations of diffusion in such environments. Vortices can be produced by applying pressure, temperature and electric potential <span class="hlt">gradients</span> via miniaturised actuators. However, the difficulties associated with the fabrication, integration, maintenance and operation of such actuators hinder their utility. Here, we utilise liquid metal enabled pumps to induce vortices inside a miniaturised liquid chamber. The configuration and rotational <span class="hlt">velocity</span> of these vortices can be controlled by tuning the polarity and frequency of the energising electrical signal. This allows creation of customised spatial temperature <span class="hlt">gradients</span> inside the chamber. The absence of conventional moving elements in the pumps facilitates the rapid reconfiguration of vortices. This enables quick transition from one temperature profile to another, and creates customised spatiotemporal temperature <span class="hlt">gradients</span>. This allows temperature oscillation from 35 to 62 °C at the hot spot, and from 25 to 27 °C at the centre of the vortex within 15 seconds. Our liquid metal enabled vortex generator can be fabricated, integrated and operated easily, and offers opportunities for studying thermo-responsive materials and biological samples.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>